Download as pdf or txt
Download as pdf or txt
You are on page 1of 385

Commutative Algebra

Pete L. Clark
Contents

Introduction 9
1. What is Commutative Algebra? 9
2. Why study Commutative Algebra? 9
3. What distinguishes this text 10
4. More on the contents 11
5. Acknowledgments 12

Chapter 1. Commutative rings 13


1. Fixing terminology 13
2. Adjoining elements 16
3. Ideals and quotient rings 17
4. The monoid of ideals of R 20
5. Pushing and pulling ideals 21
6. Maximal and prime ideals 21
7. Products of rings 22
8. A cheatsheet 24

Chapter 2. Galois Connections 27


1. The basic formalism 27
2. Lattice properties 29
3. Examples of Antitone Galois Connections 29
4. Antitone Galois Connections Decorticated: Relations 31
5. Isotone Galois Connections 32
6. Examples of Isotone Galois Connections 33

Chapter 3. Modules 35
1. Basic definitions 35
2. Finitely presented modules 40
3. Torsion and torsionfree modules 42
4. Tensor and Hom 43
5. Projective modules 45
6. Injective modules 52
7. Flat modules 59
8. Nakayama’s Lemma 61
9. Ordinal Filtrations and Applications 65
10. Tor and Ext 73
11. More on flat modules 81
12. Faithful flatness 86

Chapter 4. First Properties of Ideals in a Commutative Ring 91


3
4 CONTENTS

1. Introducing maximal and prime ideals 91


2. Radicals 93
3. Comaximal ideals 98
4. Local rings 101
5. The Prime Ideal Principle of Lam and Reyes 101
6. Minimal Primes 106
7. An application to unit groups 108

Chapter 5. Examples of Rings 109


1. Rings of numbers 109
2. Rings of continuous functions 110
3. Rings of holomorphic functions 116
4. Kapovich’s Theorem 118
5. Polynomial rings 122
6. Semigroup algebras 124

Chapter 6. Swan’s Theorem 131


1. Introduction to (topological) vector bundles 131
2. Swan’s Theorem 132
3. Proof of Swan’s Theorem 133
4. Applications of Swan’s Theorem 137
5. Stably free modules 137

Chapter 7. Localization 145


1. Definition and first properties 145
2. Pushing and pulling via a localization map 147
3. The fibers of a morphism 149
4. Commutativity of localization and passage to a quotient 149
5. Localization at a prime ideal 150
6. Localization of modules 150
7. Local properties 152
8. Local characterization of finitely generated projective modules 157

Chapter 8. Noetherian rings 161


1. Chain conditions on partially ordered sets 161
2. Chain conditions on modules 162
3. Semisimple modules and rings 163
4. Normal Series 166
5. The Krull-Schmidt Theorem 167
6. Some important terminology 171
7. Introducing Noetherian rings 172
8. Theorems of Eakin-Nagata, Formanek and Jothilingam 173
9. The Bass-Papp Theorem 176
10. Artinian rings: structure theory 177
11. The Hilbert Basis Theorem 180
12. The Krull Intersection Theorem 181
13. Krull’s Principal Ideal Theorem 184
14. The Dimension Theorem, following [BMRH] 186
15. The Artin-Tate Lemma 186
CONTENTS 5

Chapter 9. Boolean rings 189


1. First Properties 189
2. Boolean Algebras 189
3. Ideal Theory in Boolean Rings 192
4. The Stone Representation Theorem 194
5. Boolean Spaces 195
6. Stone Duality 197

Chapter 10. Associated Primes and Primary Decomposition 199


1. Associated Primes 199
2. The support of a module 202
3. Primary Ideals 203
4. Primary Decomposition, Lasker and Noether 205
5. Irredundant primary decompositions 206
6. Uniqueness properties of primary decomposition 207
7. Applications in dimension zero 210
8. Applications in dimension one 210

Chapter 11. Nullstellensätze 211


1. Zariski’s Lemma 211
2. Hilbert’s Nullstellensatz 212
3. The Real Nullstellensatz 216
4. The Combinatorial Nullstellensatz 219
5. The Finite Field Nullstellensatz 221
6. Terjanian’s Homogeneous p-Nullstellensatz 222

Chapter 12. Goldman domains and Hilbert-Jacobson rings 227


1. Goldman domains 227
2. Hilbert rings 230
3. Jacobson Rings 231
4. Hilbert-Jacobson Rings 232
5. Application: Zero-Dimensional Ideals in Polynomial Rings 233

Chapter 13. Spec R as a topological space 237


1. The Prime Spectrum 237
2. Properties of the spectrum: quasi-compactness 238
3. Properties of the spectrum: connectedness 239
4. Properties of the spectrum: separation and specialization 239
5. Irreducible spaces 242
6. Noetherianity 244
7. Krull Dimension of Topological Spaces 248
8. Jacobson spaces 248
9. Hochster’s Theorem 252
10. Rank functions revisited 253
11. The Forster-Swan Theorem 254

Chapter 14. Integral Extensions 257


1. First properties of integral extensions 257
2. Integral closure of domains 259
3. Spectral properties of integral extensions 261
6 CONTENTS

4. Integrally closed domains 262


5. The Noether Normalization Theorem 264
6. Some Classical Invariant Theory 267
7. Galois extensions of integrally closed domains 271
8. Almost Integral Extensions 272

Chapter 15. Factorization 273


1. Kaplansky’s Theorem (II) 273
2. Atomic domains, ACCP 274
3. EL-domains 276
4. GCD-domains 277
5. GCDs versus LCMs 279
6. Polynomial rings over UFDs 281
7. Application: the Schönemann-Eisenstein Criterion 285
8. Application: Determination of Spec R[t] for a PID R 286
9. Power series rings over UFDs 287
10. Nagata’s Criterion 288
11. The Euclidean Criterion 292

Chapter 16. Principal rings and Bézout domains 297


1. Principal ideal domains 297
2. Structure theory of principal rings 299
3. Euclidean functions and Euclidean rings 302
4. Bézout domains 304

Chapter 17. Valuation rings 307


1. Basic theory 307
2. Ordered commutative groups 309
3. Connections with integral closure 313
4. Another proof of Zariski’s Lemma 315
5. Discrete valuation rings 316

Chapter 18. Normalization theorems 319


1. The First Normalization Theorem 319
2. The Second Normalization Theorem 320
3. The Krull-Akizuki Theorem 321

Chapter 19. The Picard Group and the Divisor Class Group 325
1. Fractional ideals 325
2. The Ideal Closure 327
3. Invertible fractional ideals and the Picard group 328
4. Divisorial ideals and the Divisor Class Group 332

Chapter 20. Dedekind domains 335


1. Characterization in terms of invertibility of ideals 335
2. Ideal factorization in Dedekind domains 336
3. Local characterization of Dedekind domains 337
4. Factorization into primes implies Dedekind 338
5. Generation of ideals in Dedekind domains 339
6. Modules over a Dedekind domain 340
CONTENTS 7

7. Injective Modules 343

Chapter 21. Prüfer domains 345


1. Characterizations of Prüfer Domains 345
2. Butts’s Criterion for a Dedekind Domain 348
3. Modules over a Prüfer domain 350

Chapter 22. One Dimensional Noetherian Domains 351


1. Residually Finite Domains 351
2. Cohen-Kaplansky domains 354
3. Rings of Finite Rank 359
Chapter 23. Structure of overrings 365
1. Introducing overrings 365
2. Overrings of Dedekind domains 365
3. Elasticity in Replete Dedekind Domains 369
4. Overrings of Prüfer Domains 372
5. Kaplansky’s Theorem (III) 374
6. Every commutative group is a class group 375

Bibliography 379
Introduction

1. What is Commutative Algebra?


Commutative algebra is the study of commutative rings and attendant structures,
especially ideals and modules.

This is the only possible short answer I can think of, but it is not completely
satisfying. We might as well say that Hamlet, Prince of Denmark is about a fic-
tional royal family in late medieval Denmark and especially about the title (crown)
prince, whose father (i.e., the King) has recently died and whose father’s brother
has married his mother (i.e., the Queen). Informative, but not the whole story!

2. Why study Commutative Algebra?


What are the purely mathematical reasons for studying any subject of pure math-
ematics? I can think of two:

I. Commutative algebra is a necessary and/or useful prerequisite for the study


of other fields of mathematics in which we are interested.

II. We find commutative algebra to be intrinsically interesting and we want to


learn more. Perhaps we even wish to discover new results in this area.

Most beginning students of commutative algebra can relate to the first reason:
they need, or are told they need, to learn some commutative algebra for their study
of other subjects. Indeed, commutative algebra has come to occupy a remarkably
central role in modern pure mathematics, perhaps second only to category theory
in its ubiquitousness, but in a different way. Category theory provides a common
language and builds bridges between different areas of mathematics: it is something
like a circulatory system. Commutative algebra provides core results and structures
that other results and structures draw upon are overlayed upon: it is something
like a skeleton.

The branch of mathematics which most of all draws upon commutative algebra
for its structural integrity is algebraic geometry, the study of geometric properties
of manifolds and singular spaces which arise as solution sets to systems of polyno-
mial equations. There is a hard lesson here: in the 19th century algebraic geometry
split off from complex function theory and differential geometry as its own discipline
and then burgeoned dramatically at the turn of the century and the years there-
after. But by 1920 or so the practitioners of the subject had found their way into
territory in which “purely geometric” reasoning led to serious errors. In particular
9
10 INTRODUCTION

they had been making arguments about how algebraic varieties behave generically,
but they lacked the technology to even give a precise meaning to the term. Thus
the subject eventually became invertebrate and began to collapse under its own
weight. Then (starting in about 1930) came a heroic shoring up process in which
the foundations of the subject were recast with commutative algebraic methods at
the core. This was done several times over, in different ways, by Zariski, Weil, Serre
and Grothendieck, among others. For the last 60 years it has been impossible to
deeply study algebraic geometry without knowing commutative algebra – a lot of
commutative algebra. (More than is contained in these notes!)

The other branch of mathematics which draws upon commutative algebra in an


essential way is algebraic number theory. One sees this from the beginning in that
the Fundamental Theorem of Arithmetic is the assertion that the ring Z is a unique
factorization domain (UFD), a basic commutative algebraic concept. Moreover
number theory was one of the historical sources of the subject. Notably the con-
cept of Dedekind domain came from Dedekind’s number-theoretic investigations.
At the student level, algebraic number theory does not embrace commutative al-
gebra as early or as thoroughly as algebraic geometry. This seems to me to be
a pedagogical mistake: although one can do a good amount of algebraic number
theory without explicit reliance on commutative algebra, this seems to come at the
expense of not properly explaining what is going on. A modicum of commutative
algebra greatly enriches the study of algebraic number theory: it clarifies it, gener-
alizes it and (I believe) makes it more interesting.

The interplay among number theory, algebraic geometry and commutative alge-
bra flows in all directions. What Grothendieck did in the 1960s (with important
contributions from Chevalley, Serre and others) was to create a single field of math-
ematics that encompassed commutative algebra, classical algebraic geometry and
algebraic number theory: the theory of schemes. As a result, most contemporary
number theorists are also partly commutative algebraists and partly algebraic ge-
ometers: we call this cosmopolitan take on the subject arithmetic geometry.

There are other areas of mathematics that draw upon commutative algebra in
important ways. To mention some which will show up in later in these notes:

• Differential topology: vector bundles on a compact base.


• General topology.
• Invariant theory.
• Order theory.

3. What distinguishes this text


The most straightforward raison d’être for a commutative algebra text would be to
provide a foundation for the subjects of algebraic geometry, arithmetic geometry
and algebraic number theory. The bad news is that this task – even, restricted to
providing foundations for the single, seminal text of Hartshorne [Ha] – is dauntingly
large. The good news is that this has nevertheless been achieved some time ago by
David Eisenbud (a leading contemporary expert on the interface of commutative
algebra and algebraic geometry) in his text [Ei]. This work is highly recommended.
4. MORE ON THE CONTENTS 11

It is 797 pages long, so contains enough material for many courses in the subject.
It would be folly to try to improve upon (and terrible drudgery to successfully im-
itate) Eisenbud’s work here, and I certainly have not tried.

The other standard commutative algebra texts are those by Atiyah-Macdonald


[AM] and by Matsumura [M]. Any reader who is halfway serious in their study
of commutative algebra should also have access to all of [AM], [M] and [Ei] and
consult them frequently. While the current text does not rely on them in the logical
sense, I am – at best – a part time commutative algebraist, and much of what I
know comes from these texts. (Reading only “derivative” sources is rarely a good
idea.) On the other hand, precisely because there are three standard excellent texts
I have allowed my choice of topics to be, in places, much less standard.

The topics covered in Atiyah-Macdonald’s text in particular have become a de


facto standard for a first course in commutative algebra. Here are the chapter
titles from [AM]: 1. Rings and Ideals 2. Modules 3. Rings and Modules of Frac-
tions 4. Primary Decomposition 5. Integral Dependence and Valuations 6. Chain
Conditions 7. Noetherian Rings 8. Artin Rings 9. Discrete Valuation Rings and
Dedekind Domains 10. Completions 11. Dimension Theory. The entire text is 126
pages, and a distinctive feature is that a substantial portion of the text is devoted
to exercises, making [AM] one of the most amenable to student study graduate
level mathematics texts I have ever seen. The exercises are especially attractive:
some of them are easy, some of them are very challenging, and they treat both core
topics and interesting side attractions.

Much of the present text covers the material of the first nine chapters of [AM]
but with a much more leisurely, detailed exposition. Many exercises in [AM] ap-
pear as proved results here. To give a specific example, Boolean rings appear in
Exercises 1.11, 1.23, 1.24, 1.25, 3.28 of [AM], in which in particular proofs of the
Stone Representation Theorem and Stone Duality Theorems are sketched. In this
text §9 is devoted to Boolean rings, including proofs of these two results.

The failure to cover completions and basic dimension theory in this text
would be unforgivable were it not the case that [AM] covers it so nicely.

Let us also compare to [M]. Here we treat the material of the first 12 sections
of [M] except §8 (Completion and the Artin-Rees Lemma) as well as some material
from §20 (UFDs). This is less than one third of the material covered in [M].

4. More on the contents


As mentioned above, one of the distinguishing features of this text – and one of the
things which makes it much lengthier compared to the portions of [AM] and [M]
that cover the same material – is that we digress to include many “applications” to
other parts of mathematics. At one point I had the idea that every section of “core
material” should be followed by a section giving applications. This conceit was not
really feasible, but there are still some entire sections devoted to applications (gen-
erally characterized by making contact with topics outside of commutative algebra
by their relative independence from the rest of the text):
12 INTRODUCTION

• §2 on Galois connections.
• §6 on vector bundles and Swan’s Theorem.
• §9 on Boolean rings, Boolean spaces and Stone Duality.

As for significant parts of sections, we have:

• §5.2 on rings of continuous functions.


• §5.3 on rings of holomorphic functions.
• §11.3 on the real Nullstellensatz
• §11.4 on the combinatorial Nullstellensatz
• §11.5 on the finite field Nullstellensatz
• §11.6 on Terjanian’s Nullstellensatz
• §13.6 on Hochster’s Theorem
• §14.6 on invariant theory and the Shephard-Todd-Chevalley Theorem

5. Acknowledgments
Thanks to Kasper Andersen, Pablo Barenbaum, Max Bender, Martin Branden-
burg, Tom Church, John Doyle, Georges Elencwajg, Amelia Ernst, Tyler Genao,
Ernest Guico, Emil Jerabek, Jason Joseph, Keenan Kidwell, David Krumm, Allan
Lacy, Casey LaRue, Stacy Musgrave, Kedar Nadkarni, Hans Parshall, Alon Regev,
Tomasz Rzepecki, Frederick Saia, Jacob Schlather, Jack Schmidt, Mariano Suárez-
Álvarez, James Taylor, Peter Tamaroff, Matthé van der Lee and Lori D. Watson
for catching errors1 and making other useful suggestions.

1Of which many, many remain: your name could go here!


CHAPTER 1

Commutative rings

1. Fixing terminology
We are interested in studying properties of commutative rings with unity.

By a general algebra R, we mean a triple (R, +, ·) where R is a set endowed


with a binary operation + : R × R → R – called addition – and a binary operation
· : R × R → R – called multiplication – satisfying the following:

(CG) (R, +) is a commutative group,

(D) For all a, b, c ∈ R, (a + b) · c = a · c + b · c, a · (b + c) = a · b + a · c.

For at least fifty years, there has been agreement that in order for an algebra
to be a ring, it must satisfy the additional axiom of associativity of multiplication:

(AM) For all a, b, c ∈ R, a · (b · c) = (a · b) · c.

A general algebra which satisfies (AM) will be called simply an algebra. A similar
convention that is prevalent in the literature is the use of the term nonassociative
algebra to mean what we have called a general algebra: i.e., a not necessarily
associative algebra.

A ring R is said to be with unity if there exists a multiplicative identity, i.e.,


an element e of R such that for all a ∈ R we have e · a = a · e = a. If e and e0
are two such elements, then e = e · e0 = e0 . In other words, if a unity exists, it is
unique, and we will denote it by 1.

A ring R is commutative if for all x, y ∈ R, x · y = y · x.

In these notes we will be (almost) always working in the category of commuta-


tive rings with unity. In a sense which will shortly be made precise, this means
that the identity 1 is regarded as part of the structure of a ring and must therefore
be preserved by all homomorphisms.

Probably it would be more natural to study the class of possibly non-commutative


rings with unity, since, as we will see, many of the fundamental constructions of
rings give rise, in general, to non-commutative rings. But if the restriction to
commutative rings (with unity!) is an artifice, it is a very useful one, since two
of the most fundamental notions in the theory, that of ideal and module, become
13
14 1. COMMUTATIVE RINGS

significantly different and more complicated in the non-commutative case. It is


nevertheless true that many individual results have simple analogues in the non-
commutative case. But it does not seem necessary to carry along the extra general-
ity of non-commutative rings; rather, when one is interested in the non-commutative
case, one can simply remark “Proposition X.Y holds for (left) R-modules over a
noncommutative ring R.”

Notation: Generally we shall abbreviate x · y to xy. Moreover, we usually do


not use different symbols to denote the operations of addition and multiplication
in different rings: it will be seen that this leads to simplicity rather than confusion.

Group of units: Let R be a ring with unity. An element x ∈ R is said to be a


unit if there exists an element y such that xy = yx = 1.

convention on exercises: Throughout the exercises, a “ring” means a com-


mutative ring unless explicit mention is made to the contrary. Some but not all of
the results in the exercises still hold for non-commutative rings, and it is left to the
interested reader to explore this.
Exercise 1.1. a) Sho: if x is a unit, the element y with xy = yx = 1 is unique,
denoted x−1 .
b) Show: if x is a unit, so is x−1 .
c) Show that, for all x, y ∈ R, xy is a unit ⇐⇒ x and y are both units.
d) Show: the units form a commutative group, denoted R× , under multiplication.
Remark 1. For elements x, y in a non-commutative ring R, if x and y are units
so is xy, but the converse need not hold. (Thus Exercise 1.1c) is an instance of a
result in which commutativity is essential.) Nevertheless this is enough to deduce
that in any ring the units R× form a group...which is not necessarily commutative.
Example 1.1 (Zero ring). Our rings come with two distinguished elements, the
additive identity 0 and the multiplicative identity 1. Suppose that 0 = 1. Then for
x ∈ R, x = 1 · x = 0 · x, whereas in any ring 0 · x = (0 + 0) · x = 0 · x + 0 · x, so
0 · x = 0. In other words, if 0 = 1, then this is the only element in the ring. It is
clear that for any one element set R = {0}, 0 + 0 = 0 · 0 = 0 endows R with the
structure of a ring. We call this ring the zero ring.
The zero ring exhibits some strange behavior, such that it must be explicitly ex-
cluded in many results. For instance, the zero element is a unit in the zero ring,
which is obviously not the case in any nonzero ring. A nonzero ring in which every
nonzero element is a unit is called a division ring. A commutative division ring
is called a field.

Let R and S be rings. A homomorphism f : R → S is a function such that:

(HOM1) For all x, y ∈ R, f (x + y) = f (x) + f (y).


(HOM2) For all x, y ∈ R, f (xy) = f (x)f (y).
(HOM3) f (1) = 1.

We observe that (HOM1) implies f (0) = f (0 + 0) = f (0) + f (0), so f (0) = 0.


Thus we do not need to explcitly include f (0) = 0 in the definition of a group
1. FIXING TERMINOLOGY 15

homomorphism. For the multiplicative identity however, this argument only shows
that if f (1) is a unit, then f (1) = 1. Therefore, if we did not require (HOM3), then
for instance the map f : R → R, f (x) = 0 for all x, would be a homomorphism,
and we do not want this.
Exercise 1.2. Suppose R and S are rings, and let f : R → S be a map
satisfying (HOM1) and (HOM2). Show that f is a homomorphism of rings (i.e.,
satisfies also f (1) = 1) iff f (1) ∈ S × .
A homomorphism f : R → S is an isomorphism if there exists a homomorphism
g : S → R such that: for all x ∈ R, g(f (x)) = x; and for all y ∈ S, f (g(y)) = y.
Exercise 1.3. Let f : R → S be a homomorphism of rings. Show the following
are equivalent:
(i) f is a bijection.
(ii) f is an isomorphism.
In many algebra texts, an isomorphism of rings (or groups, etc.) is defined to be
a bijective homomorphism, but this gives the wrong idea of what an isomorphism
should be in other mathematical contexts (e.g. for topological spaces). Rather,
having defined the notion of a morphism of any kind, one defines isomorphism in
the way we have above.
Exercise 1.4. a) Suppose R and S are both rings on a set containing exactly
one element. Show that there is a unique ring isomorphism from R to S. (This is
a triviality, but explains why are we able to speak of the zero ring, rather than
simply the zero ring associated to one element set. We will therefore denote the
zero ring just by 0.)
b) Show that any ring R admits a unique homomorphism to the zero ring. One
says that the zero ring is the final object in the category of rings.
Exercise 1.5. Show: for a not-necessarily-commutative-ring S there exists a
unique homomorphism from the ring Z of integers to S. (Thus Z is the initial
object in the category of not-necessarily-commutative-rings. It follows immediately
that it is also the initial object in the category of rings.)
A subring R of a ring S is a subset R of S such that

(SR1) 1 ∈ R.
(SR2) For all r, s ∈ R, r + s ∈ R, r − s ∈ R, and rs ∈ R.

Here (SR2) expresses that the subset R is an algebra under the operations of addi-
tion and multiplication defined on S. Working, as we are, with rings with unity, we
have to be a bit more careful: in the presence of (SR2) but not (SR1) it is possible
that R either does not have a multiplicative identity or, more subtly, that it has a
multiplicative identity which is not the element 1 ∈ S.

An example of the first phenomenon is S = Z, R = 2Z. An example of the


second is S = Z, R = 0. A more interesting example is S = Z × Z – i.e., the set
of all ordered pairs (x, y), x, y ∈ Z with (x1 , y1 ) + (x2 , y2 ) = (x1 + x2 , y1 + y2 ),
(x1 , y1 ) · (x2 , y2 ) = (x1 x2 , y1 y2 ) – and R = {(0, y) | y ∈ Z}. Then with the induced
addition and multiplication from S, R is isomorphic to the ring Z and the element
16 1. COMMUTATIVE RINGS

(0, 1) serves as a multiplicative identity on R which is different from the (always


unique) multiplicative identity 1S = (1, 1), so according to our conventions R is not
a subring of S.

Notice that if R is a subring of S, the inclusion map R ,→ S is an injective homo-


morphism of rings. Conversely, if ι : R ,→ S is an injective ring homomorphism,
then R ∼ = ι(R) and ι(R) is a subring of S, so essentially we may use ι to view R
as a subring of S. The only proviso here is that this certainly depends on ι: in
general there may be other injective homomorphisms ι : R ,→ S which realize R as
a different subset of S, hence a different subring.

2. Adjoining elements
Let ι : R ,→ S be an injective ring homomorphism. As above, let us use ι to view
R as a subring of S; we also say that S is an extension ring of R and write S/R
for this (note: this has nothing to do with cosets or quotients!) We wish now to
consider rings T such that R ⊂ T ⊂ S; such a ring T might be called a subexten-
sion of S/R or an intermediate ring.

For ι : R ,→ S as above, let X = {xi } be a subset of S. Then the partially


ordered set of all subrings of S containing R and X is nonempty (since S is in it)
and contains a bottom element, given (as usual!) by taking the intersection of all
of its elements. We call this the ring obtained by adjoining the elements of X to
R. In the commutative case, we denote this ring by R[{xi }], for reasons that will
become more clear when we discuss polynomial rings in §5.4.

Example 1.2. Take R = Z,√ S = C. Then Z[i] = Z[ −1] is the smallest
subring of C containing (Z and) −1.
Example 1.3. Take R = Z, S = Q, let P be any set of prime numbers, and
put X = { p1 }p∈P . Then there is a subring ZP := Z[{ p1 }p∈P ] of Q.
Exercise 1.6. Let P, Q be two sets of prime numbers. Show the following are
equivalent:
(i) ZP ∼= ZQ .
(ii) ZP = ZQ .
(iii) P = Q.
Exercise 1.7. Show: every subring of Q is of the form ZP for some P.
The adjunction process R 7→ R[X] is defined only relative to some extension ring S
of R, although the notation hides this. In fact, one of the recurrent themes of the
subject is the expression of the adjunction process in a way which
√ depends only on
R itself. In the first example, this is√achieved by identifying −1 with its minimal
polynomial t2 + 1 and replacing Z[ −1] with the quotient ring Z[t]/(t2 + 1). The
second example will eventually be turned around: we will be able to give an inde-
pendent definition of ZP as a certain “ring of fractions” formed from Z and then Q
will be the ring of fractions obtained by taking P to be the set of all prime numbers.

Nevertheless, the existence of such turnabouts should not cause us to forget that
adjunction is relative to √an extension; indeed forgetting this can lead to serious
trouble. For instance, if 3 2 is the unique real cube root of 2 and ζ3 is a primitive
3. IDEALS AND QUOTIENT RINGS 17


cube √ root of unity, then
√ the three complex numbers with cube 2 are z1 = 3 2,
z2 = 3 2ζ3 and z3 = 3 2ζ32 . Each of the rings Q[z1 ], Q[z2 ], Q[z3 ] is isomorphic to
the ring Q[t]/(t3 − 2), so all three are isomorphic to each other. But they are not
the same ring: on the one hand Q[z √ 1 ] is contained in R and the other two are not.
More seriously Q[z1 , z2 , z3 ] = Q[ 3 2, ζ3 ], which strictly contains any one of Q[z1 ],
Q[z2 ] and Q[z3 ].

3. Ideals and quotient rings


Let f : R → S be a homomorphism of rings, and put
I = f −1 (0) = {x ∈ R | f (x) = 0}.
In particular f is a homomorphism of commutative groups (R, +) → (S, +), I is a
subgroup of (R, +). Moreover, it enjoys both of the following properties:

(LI) For all j ∈ I and x ∈ R, xj ∈ I.


(RI) For all i ∈ I and y ∈ R, iy ∈ I.

Indeed,
f (xj) = f (x)f (j) = f (x) · 0 = 0 = 0 · f (y) = f (i)f (y) = f (iy).
In general, let R be a ring. An ideal is a subset I ⊂ R which is a subgroup of
(R, +) (in particular, 0 ∈ I) and which satisfies (LI) and (RI).
Theorem 1.4. Let R be a ring, and let I be a subgroup of (R, +). The following
are equivalent:
(i) I is an ideal of R.
(ii) There exists a ring structure on the quotient group R/I making the additive
homomorphism R → R/I into a homomorphism of rings.
When these conditions hold, the ring structure on R/I in (ii) is unique, and R/I
is called the quotient of R by the ideal I.
Proof. Consider the group homomorphism q : R → R/I. If we wish R/I to
be a ring in such a way so that q is a ring homomorphism, we need
(x + I)(y + I) = q(x)q(y) = q(xy) = (xy + I).
This shows that there is only one possible ring structure, and the only question is
whether it is well-defined. For this we need that for all i, j ∈ I, (x + i)(y + j) − xy =
xj + iy + ij ∈ I. Evidently this holds for all x, y, i, j iff (LI) and (RI) both hold. 
Remark: If R is commutative, then of course there is no difference between (LI) and
(RI). For a non-commutative ring R, an additive subgroup I satisfying condition
(LI) but not necessarily (RI) (resp. (RI) but not necessarily (LI)) is called a left
ideal (resp. a right ideal). Often one says two-sided ideal to emphasize that
(LI) and (RI) both hold. Much of the additional complexity of the non-commutative
theory comes from the need to distinguish between left, right and two-sided ideals.

We do not wish to discuss such complexities here, so henceforth in this section


we assume (except in exercises, when indicated) that our rings are commutative.

Example: In R = Z, for any integer n, consider the subset (n) = nZ = {nx | x ∈ Z}


18 1. COMMUTATIVE RINGS

of all multiples of n. This is easily seen to be an ideal.1 The quotient Z/nZ is the
ring of integers modulo n.

An ideal I ( R is called proper.


Exercise 1.8. Let R be a ring and I an ideal of R. Show the following are
equivalent:
(i) I ∩ R× 6= ∅.
(ii) I = R.
Exercise 1.9. a) Let R be a commutative ring. Show that R is a field iff R
has exactly two ideals, 0 and R.
b) Let R be a not necessarily commutative ring. Show the following are equivalent:
(i) The only one-sided ideals of R are 0 and R. (ii) R is a division ring.
c) For a field k and an integer n > 1, show that the matrix ring Mn (k) has no
two-sided ideals but is not a division ring.
Exercise 1.10. Some contemporary undergraduate algebra texts define the fi-
nite ring Z/nZ in a different and apparently simpler way: put Zn = {0, 1, . . . , n−1}.
For any integer x, there is a unique integer k such that x − kn ∈ Zn . Define a func-
tion mod n : Z → Zn by mod n(x) := x − kn. We then define + and · on Zn by
x + y := mod n(x + y), xy = mod n(xy). Thus we have avoided any mention of
ideals, equivalence classes, quotients, etc. Is this actually simpler? (Hint: how do
we know that Zn satisfies the ring axioms?)
For any commutative ring R and any element y ∈ R, the subset (y) = yR =
{xy | x ∈ R} is an ideal of R. Such ideals are called principal. A principal ideal
ring is a commutative ring in which each ideal is principal.
Exercise 1.11. a) The intersection of any family of (left, right or two-sided)
ideals in a not-necessarily-commutative-ring is a (left, right or two-sided)
T ideal.
b) Let {Ii } be a set of ideals in the commutative ring R. Show that T i Ii has the
following property: for any ideal J of R such that J ⊂ Ii for all i, J ⊂ i I.
Let R be a ring
T and S a subset of R. There is then a smallest ideal of R containing
S, namely Ii , where Ii are all the ideals of R containing S. We call this the
ideal generated by S. This is a “top-down” description; as usual, there is a
complementary “bottom-up” description which is not quite as clean but often more
useful. Namely, put X
hSi := { ri si | ri ∈ R, si ∈ S}
i.e., the set of all finite sums of an element of R times an element of S.
Proposition 1.5. For a subset S of a commutative ring R, hSi is an ideal,
the intersection of all ideals containing S.
Exercise 1.12. Prove Proposition 1.5.
When S is a subset of R such that I = hSi, we say S is a set of generators for
I. In general the same ideal will have many (most often infinitely many) sets of
generators.

1If this is not known and/or obvious to the reader, these notes will probably be too brisk.
3. IDEALS AND QUOTIENT RINGS 19

An ideal I is principal if it can be generated by a single element. In any ring,


the zero ideal 0 = h0i and the entire ring R = h1i are principal. For x ∈ R, we
tend to denote the principal ideal generated by x as either Rx or (x) rather than hxi.

An ideal I is finitely generated if...it admits a finite set of generators.2

Stop and think for a moment: do you know an example of an ideal which is not
finitely generated? You may well find that you do not. It turns out that there is a
very large class of rings – including most or all of the rings you are likely to meet
in undergraduate algebra – for which every ideal is finitely generated. A ring R
in which every ideal is finitely generated is called Noetherian. This is probably
the single most important class of rings, as we will come to appreciate slowly but
surely over the course of these notes.
Exercise 1.13. Let R be a ring.
a) For ideals I and J of R, define I + J = {i + j | i ∈ I, j ∈ J}. Show that
I + J = hI ∪ Ji is the smallest ideal containing both I and J.
b) Extend part a) to any finite number of ideals I1 , . . . , In .
c) Suppose {Ii } is a set of ideals of I. Give an explicit description of the ideal hIi i.
The preceding considerations show that the collection of all ideals of a commutative
ring R, partially ordered by inclusion, form a complete lattice.

If I is an ideal in the ring R, then there is a correspondence between ideals J


of R containing I and ideals of the quotient ring R/I, exactly as in the case of a
normal subgroup of a group:
Theorem 1.6. (Correspondence theorem) Let I be an ideal of a ring R, and
denote the quotient map R → R/I by q. Let I(R) be the lattice of ideals of R,
II (R) be the sublattice of ideals containing I and I(R/I) the lattice of ideals of the
quotient ring R/I. Define maps
Φ : I(R) → I(R/I), J 7→ (I + J)/I,
Ψ : I(R/I) → I(R), J 7→ q −1 (J).
Then Ψ ◦ Φ(J) = I + J and Φ ◦ Ψ(J) = J. In particular Ψ induces an isomorphism
of lattices from I(R/I) to II (R).
Proof. For all the abstraction, the proof is almost trivial. For J ∈ I(R), we
check that Ψ(Φ(J)) = Ψ(J +I (mod I)) = {x ∈ R | x+I ∈ J +I} = J +I ∈ II (R).
Similarly, for J ∈ I(R/I), we have Φ(Ψ(J)) = J. 
Remark: In fancier language, the pair (Φ, Ψ) give an isotone Galois connection
between the partially ordered sets I(R) and I(R/I). The associated closure oper-
ator Φ ◦ Ψ on I(R/I) is the identity, whereas the closure operator Ψ ◦ Φ on I(R)
carries each ideal J to the smallest ideal containing both J and I.3

The Correspondence Theorem will be our constant companion. As is common,


we will often use the map Ψ to identify the sets I(R/I) and II (R).
2Well, obviously. Nevertheless this definition is so critically important that it would have
been a disservice to omit it.
3This point of view will be explored in more detail in §2.
20 1. COMMUTATIVE RINGS

Exercise 1.14. Let I be an ideal of R and {Ji } be a set of ideals of R. Show


that Φ preserves suprema and Ψ preserves infima:
Φ(hJi i) = hΦ(Ji )i
and \ \
Ψ( Ji ) = Ψ(Ji ).

4. The monoid of ideals of R


Let I and J be ideals of the ring R. The product ideal IJ is the least ideal
P elements of the form xy for x ∈ I and y ∈ J. (It is easy to see
containing all
that IJ = { xi yi | xi ∈ I, yi ∈ J} is precisely the set of all finite sums of such
products.) Recall that we have written I(R) for the lattice of all ideals of R. Then
(I, J) 7→ IJ gives a binary operation on I(R), the ideal product.
Exercise 1.15. Show that I(R) under the ideal product is a commutative
monoid, with identity element R and absorbing element the (0) ideal of R.4
If you are given a commutative monoid M , then invariably the property you are
hoping it has is cancellation: for all x, y, z ∈ M , xz = yz =⇒ x = y.5 For
example, if R is a ring, then the set R• of nonzero elements of R is cancellative iff
R is a domain. In (R, ·) 0 is an absorbing element, so we remove it to get a hope
of cancellativity.
Exercise 1.16. a) Let M be a cancellative monoid of cardinality greater than
one. Show that M does not have any absorbing elements.
b) Let R be a ring which is not the zero ring. Show that the monoid I(R) is not
cancellative.
In light of the previous exercise, for a domain R we define I • (R) to be the monoid
of nonzero ideals of R under multiplication.

Warning: Just because R is a domain, I • (R) need not be cancellative!


√ √ √
Exercise 1.17. Let R = Z[ −3], and let p2 = h1 + −3, 1 − −3i (i.e., the
ideal generated by these two elements.
a) Show that #R/(2) = 4 and R/p2 ∼ = Z/2Z.
b) Show that p22 = p2 · (2).
c) Conclude that I • (R) is not cancellative.
Exercise 1.18. Let R be a PID. Show that I • (R) is cancellative.
Exercise 1.19. Show: for a commutative monoid M , the following are equiv-
alent:
(i) M is cancellative.
(ii) There exists a commutative group G and an injective monoid homomorphism
ι : M ,→ G.
Exercise 1.20. Let M be a commutative monoid. A group completion of
M consists of a commutative group G(M ) and a monoid homomorphism F : M →
G(M ) which is universal for monoid homomorphisms into a commutative group.
4An element z of a monoid M is called absorbing if for all x ∈ M , zx = xz = z.
5Well, obviously this is an exaggeration, but you would be surprised how often it is true.
6. MAXIMAL AND PRIME IDEALS 21

That is, for any commutative group G and monoid homomorphism f : M → G,


there exists a unique homomorphism of groups q : G → G(M ) such that F = q ◦ f .
a) Show that any two group completions are isomorphic.
b) Show that any commutative monoid has a group completion.
c) Show that a commutative monoid injects into its group completion iff it is can-
cellative.

5. Pushing and pulling ideals


Let f : R → S be a homomorphism of commutative rings. We can use f to trans-
port ideals from R to S and also to transport ideals from S to R.

More precisely, for I an ideal of R, consider f (I) as a subset of S.


Exercise 1.21. a) Give an example to show that f (I) need not be an ideal of
S.
b) Suppose f is surjective. Show: f (I) is an ideal of S.
Nevertheless we can consider the ideal it generates: we define
f∗ (I) = hf (I)i,
and we call f∗ (I) the pushforward of I to S.

Similarly, let J be an ideal of S, and consider its complete preimage in R, i.e.,


f −1 (J) = {x ∈ R | f (x) ∈ J}. As you are probably already aware, preimages have
much nicer algebraic properties than direct images, and indeed f −1 (J) is necessar-
ily an ideal of R. We denote it by f ∗ (J) and call it the pullback of J to R.

Example: Suppose that I is an ideal of R, S = R/I and f : R → R/I is the


quotient map. In this case, pushforwards and pullbacks were studied in detail in
Theorem 1.6. In this case f ∗ : I(S) ,→ I(R) is an injection, which allows us to
view the lattice of ideals of S as a sublattice of the lattice of ideals of R. Moreover
we have a push-pull formula: for all ideals J of R,
f ∗ f∗ J = J + I
and also a pull-push formula: for all ideals J of R/I,
f∗ f ∗ J = J.
These formulas are extremely useful at all points in the study of ring theory. More
generally, whenever one meets a homomorphism f : R → S of rings (or better, a
certain class of homomorphisms), it is fruitful to ask about properties of f∗ and
f ∗ : in particular, is f ∗ necessarily injective, or surjective? Can we identify the
composite maps f ∗ f∗ and/or f∗ f ∗ ?
In these notes, the most satisfying and important answers will come for local-
izations and integral extensions.

6. Maximal and prime ideals


An ideal m of R is maximal if it is proper and there is no proper ideal of R strictly
containing m. An ideal p of R is prime if for all x, y ∈ R, xy ∈ p implies x ∈ p or
y ∈ p or both.
22 1. COMMUTATIVE RINGS

Exercise 1.22. For an ideal I in a ring R, show: the following are equivalent:
(i) I is maximal.
(ii) R/I is a field.
Exercise 1.23. For an ideal I in a ring R, show: the following are equivalent:
(i) I is prime.
(ii) R/p is a domain.
Exercise 1.24. Show: maximal ideals are prime.
Exercise 1.25. Let f : R → S be a homomorphism of rings.
a) Let I be a prime ideal of R. Show that f∗ I need not be a prime ideal of S.
b) Let J be a prime ideal of S. Show that f ∗ J is a prime ideal of R.
c) Let J be a maximal ideal of S. Show that f ∗ J need not be maximal in R.
If I and J are ideals of a ring R, we define the colon ideal6
(I : J) = {x ∈ R | xJ ⊂ I}.
Exercise 1.26. Show: (I : J) is indeed an ideal of R.

7. Products of rings
Let R1 and R2 be rings. The Cartesian product R1 × R2 has the structure of a ring
with “componentwise” addition and multiplication:

(r1 , r2 ) + (s1 , s2 ) := (r1 + s1 , r2 + s2 ).


(r1 , r2 ) · (s1 , s2 ) := (r1 s1 , r2 s2 ).
Exercise 1.27. a) Show that R1 × R2 is commutative iff both R1 and R2 are
commutative.
b) R1 × R2 has an identity iff both R1 and R2 do, in which case e := (e1 , e2 ) is the
identity of R1 × R2 .
As for any Cartesian product, R1 × R2 comes equipped with its projections
π1 : R1 × R2 → R1 , | (r1 , r2 ) 7→ r1
π2 : R1 × R2 → R2 , | (r1 , r2 ) 7→ r2 .
The Cartesian product X1 × X2 of sets X1 and X2 satisfies the following universal
property: for any set Z and any maps f1 : Z → X1 , f2 : Z → X2 , there exists a
unique map f : Z → X1 × X2 such that f1 = π1 ◦ f , f2 = π2 ◦ f . The Cartesian
product R1 × R2 satisfies the analogous universal property in the category of rings.
Exercise 1.28. For rings R1 , R2 , S and ring homomorphisms fi : S → Ri ,
there exists a unique homomorphism of rings f : S → R1 × R2 such that fi = πi ◦ f .
So the Cartesian product of R1 and R2 is also the product in the categorical sense.

As with sets, we can equally well take the Cartesian product over an arbitrary
indexed
Q family of rings: if {Ri }i∈I is a family of rings, their Cartesian product
i∈I R i becomes a ring under coordinatewise addition and multiplication, and sat-
isfies the universal property of the product. Details are left to the reader.
6The terminology is unpleasant and is generally avoided as much as possible. One should
think of (I : J) as being something like the “ideal quotient” I/J (which of course has no formal
meaning). Its uses will gradually become clear.
7. PRODUCTS OF RINGS 23

It is natural to ask whether the category of rings has a direct sum as well. In
other words, given rings R1 and R2 we are looking for a ring R together with
ring homomorphisms ιi : Ri → R such that for any ring S and homomorphisms
fi : Ri → S, there exists a unique homomorphism f : R → S such that fi = f ◦ ιi .
We recall that in the category of commutative groups, the Cartesian product group
G1 × G2 also the categorical direct sum, with ι1 : g 7→ (g, 0) and ι2 : g 7→ (0, g).
Since each ring has in particular the structure of a commutative group, it is nat-
ural to wonder whether the same might hold true for rings. However, the map
ι1 : R1 → R1 × R2 does not preserve the multiplicative identity (unless R2 = 0),
so is not a homomorphism of rings when identities are present. Moreover, even
in the category of algebras, in order to satisfy the universal property on the un-
derlying additive subgroups, the homomorphism f is uniquely determined to be
(r1 , r2 ) 7→ f1 (r1 ) + f2 (r2 ), and it is easily checked that this generally does not pre-
serve the product.

(The category of rings does have direct sums in the categorical sense: the cate-
gorical direct sum of R1 and R2 is given by the tensor product R1 ⊗Z R2 .)

Now returning to the case of commutative rings, let us consider the ideal structure
of the product R = R1 × R2 . If I1 is an ideal of R1 , then I1 × {0} = {(i, 0) | i ∈ I}
is an ideal of the product; moreover the quotient R/(I1 × {0} is isomorphic to
R1 /I1 × R2 . Similarly, if I2 is an ideal, {0} × I2 is an ideal of R2 . Finally, if I1 is
an ideal of R1 and I2 is an ideal of R2 , then
I1 × I2 := {(i1 , i2 ) |i1 ∈ I1 , i2 ∈ I2 }
is an ideal of R. In fact we have already found all the ideals of the product ring:
Proposition 1.7. Let R1 and R2 be commutative rings, and let I be an ideal of
R := R1 × R2 . For i = 1, 2, put Ii := πi (I). Then for i = 1, 2 we have that πi (I) is
an ideal of Ri and I = π1 (I)×π2 (I). Then I = I1 ×I2 = {(i1 , i2 ) | i1 ∈ I1 , i2 ∈ I2 }.
Proof. Since πi : R1 × R2 → Ri is a surjective ring homomorphism, by
Exercise 1.21b) we have that πi (I) is an ideal of Ri , and thus π1 (I) × π2 (I) is an
ideal of R1 × R2 .
For any subset S of a Cartesian product X1 × X2 we have S ⊂ π1 (S) × π2 (S),
so certainly
I ⊂ π1 (I) × π2 (I).
The reverse inclusion certainly does not hold in general for subsets of Cartesian
products, but it does hold here: if x ∈ π1 (I) then there is (x, y) ∈ I and then
(x, 0) = (x, y) · (1, 0) ∈ I. Similarly we get that if y ∈ π2 (I) then (0, y) ∈ I, so
(x, y) = (x, 0) + (0, y) ∈ I + I = I. 
Another way to express the result is that, corresponding to a decomposition R =
R1 × R2 , we get a decomposition I(R) = I(R1 ) × I(R2 ).

Let us call a commutative ring R disconnected if there exists nonzero rings R1 ,


R2 such that R ∼ = R1 × R2 , and connected otherwise.7 If R is disconnected,
7We will see later that there is a topological space Spec R associated to every ring, and Spec R
is a disconnected topological space iff R can be written as a nontrivial product of rings
24 1. COMMUTATIVE RINGS

then choosing such an isomorphism ϕ, we may put I1 = ϕ−1 (R1 × {0}) and
I2 = ϕ−1 ({0} × R2 ). Evidently I1 and I2 are ideals of R such that I1 ∩ I2 = {0} and
I1 + I2 = R. Conversely, if in a ring R we can find a pair of ideals I1 , I2 with these
properties then it will follow from the Chinese Remainder Theorem (Theorem 4.20)
that the natural map Φ : R → R/I2 × R/I1 , r 7→ (r + I2 , r + I1 ) is an isomorphism.

Now Φ restricted to I1 induces an isomorphism of groups onto R/I2 (and similarly


with the roles of I1 and I2 reversed). We therefore have a distinguished element
of I1 , e1 := Φ−1 (1). This element e1 is an identity for the multiplication on R re-
stricted to I1 ; in particular e21 = e1 ; such an element is called an idempotent. In
any ring the elements 0 and 1 are idempotents, called trivial; since e1 = Φ−1 (1, 0) –
and not the preimage of (0, 0) or of (1, 1) – e1 is a nontrivial idempotent. Thus
a nontrivial decomposition of a ring implies the presence of nontrivial idempotents.

The converse is also true:


Proposition 1.8. Suppose R is a ring and e is a nontrivial idempotent element
of R: e2 = e but e 6= 0, 1. Put I1 = Re and I2 = R(1 − e). Then I1 and I2 are
ideals of R such that I1 ∩ I2 = 0 and R = I1 + I2 , and therefore R ∼
= R/I1 × R/I2
is a nontrivial decomposition of R.
Exercise 1.29. Prove Proposition 1.8.
Exercise 1.30. Generalize the preceding discussion to decompositions into a
finite number of factors: R = R1 × · · · × Rn .

8. A cheatsheet
Let R be a commutative ring. Here are some terms that we will analyze in lov-
ing detail later, but would like to be able to mention in passing whenever necessary.

R is a domain if xy = 0 =⇒ x = 0 or y = 0.

An ideal p of R is prime if the quotient ring R/p is a domain. Equivalently, p


is an ideal such that xy ∈ p =⇒ x ∈ p or y ∈ p.

An ideal m of R is maximal if it is proper – i.e., not R itself – and not strictly


contained in any larger proper ideal. Equivalently, m is an ideal such that the quo-
tient ring R/m is a field.

R is Noetherian if it satisfies any of the following equivalent conditions:8


(i) For any nonempty set S of ideals of R, there exists I ∈ S which is not properly
contained in any J ∈ S.
(ii) There is no infinite sequence of ideals I1 ( I2 ( . . . in R.
(iii) Every ideal of R is finitely generated.

R is Artinian (or sometimes, an Artin ring) if the partially ordered set of ideals
of R satisfies the descending chain condition: there is no infinite sequence of ideals
I1 ) I2 ) . . ..
8See Theorem 8.24 for a proof of their equivalence.
8. A CHEATSHEET 25

Let R ⊂ S be an inclusion of rings. We say that s ∈ S is integral over R if


there are a0 , . . . , an−1 ∈ R such that
sn + an−1 sn−1 + . . . + a1 s + a0 = 0.
We say that S is integral over R if every element of S is integral over R. This is
the appropriate generalization to rings of the notion of an algebraic field extension.
We will study integral elements and extensions, um, extensively in § 14, but there
is one easy result that we will need earlier, so we give it now.
Proposition 1.9. Let R ⊂ S be an integral extension of domains. If S is a
field then R is a field.
Proof. Let α ∈ R• . Then α−1 is integral over R: there are ai ∈ R such that
α−n = an−1 α−n+1 + . . . + a1 α−1 + a0 .
Multiplying through by αn−1 gives
α−1 = an−1 + an−2 α + . . . + a1 αn−2 + a0 αn−1 ∈ R. 
CHAPTER 2

Galois Connections

1. The basic formalism


Let (X, ≤) be a partially ordered set. We denote by X ∨ the order dual of X: it
has the same underlying set as X but the inverse order relation: x  y ⇐⇒ y ≤ x.

Let (X, ≤) and (Y, ≤) be partially ordered sets. A map f : X → Y is isotone


(or order-preserving) if for all x1 , x2 ∈ X, x1 ≤ x2 =⇒ f (x1 ) ≤ f (x2 ); f is
antitone (or order-reversing) if for all x1 , x2 ∈ X, x1 ≤ x2 =⇒ f (x1 ) ≥ f (x2 ).
Exercise 2.1. Let X, Y, Z be partially ordered sets, and let f : X → Y , g :
Y → Z be functions. Show:
a) If f and g are isotone, then g ◦ f is isotone.
b) If f and g are antitone, then g ◦ f is isotone.
c) If one of f and g is isotone and the other is antitone, then g ◦ f is antitone.
Let (X, ≤) and (Y, ≤) be partially ordered sets. An antitone Galois connection
between X and Y is a pair of maps Φ : X → Y and Ψ : Y → X such that:

(GC1) Φ and Ψ are both antitone maps, and


(GC2) For all x ∈ X and all y ∈ Y , x ≤ Ψ(y) ⇐⇒ y ≤ Φ(x).

There is a pleasant symmetry in the definition: if (Φ, Ψ) is a Galois connection


between X and Y , then (Ψ, Φ) is a Galois connection between Y and X.

If (X, ≤) is a partially ordered set, then a mapping f : X → X is called a closure


operator if it satisfies all of the following properties:

(C1) For all x ∈ X, x ≤ f (x).


(C2) For all x1 , x2 ∈ X, x1 ≤ x2 =⇒ f (x1 ) ≤ f (x2 ).
(C3) For all x ∈ X, f (f (x)) = f (x).
Proposition 2.1. The mapping Ψ ◦ Φ is a closure operator on (X, ≤) and the
mapping Φ ◦ Ψ is a closure operator on (Y, ≤).
Proof. By symmetry, it is enough to consider the mapping x 7→ Ψ(Φ(x)) on
X.
If x1 ≤ x2 , then since both Φ and Ψ are antitone, we have Φ(x1 ) ≥ Φ(x2 ) and
thus Ψ(Φ(x1 )) ≤ Ψ(Φ(x1 )): (C2).
For x ∈ X, Φ(x) ≥ Φ(x), and by (GC2) this implies x ≤ Ψ(Φ(x)): (C1).
Finally, for x ∈ X, applying (C1) to the element Ψ(Φ(x)) of X gives
Ψ(Φ(x)) ≤ Ψ(Φ(Ψ(Φ(x)))).
27
28 2. GALOIS CONNECTIONS

Conversely, we have
Ψ(Φ(x)) ≤ Ψ(Φ(x)),
so by (GC2)
Φ(Ψ(Φ(x)) ≥ Φ(x),
and applying the order-reversing map Ψ gives

Ψ(Φ(Ψ(Φ(x)))) ≤ Ψ(Φ(x)).

Thus
Ψ(Φ(x)) = Ψ(Φ(Ψ(Φ(x))).


Corollary 2.2. The following tridempotence properties are satisfied by Φ


and Ψ:
a) For all x ∈ X, ΦΨΦx = Φx.
b) For all y ∈ X, ΨΦΨy = Ψy.

Proof. By symmetry, it suffices to prove a). Since Φ ◦ Ψ is a closure operator,


ΦΨΦx ≥ Φx. Moreover, since Ψ ◦ Φ is a closure operator, ΨΦx ≥ x, and since Φ is
antitone, ΦΨΦx ≤ Φx. So ΦΨΦx = Φx. 

Proposition 2.3. Let (Φ, Ψ) be a Galois connection between partially ordered


sets X and Y . Let X = Ψ(Φ(X)) and Y = Ψ(Φ(Y )).
a) X and Y are precisely the subsets of closed elements of X and Y respectively.
b) We have Φ(X) ⊂ Y and Ψ(Y ) ⊂ X.
c) Φ : X → Y and Ψ : Y → X are mutually inverse bijections.

Proof. a) If x = Ψ(Φ(x)) then x ∈ X. Conversely, if x ∈ X, then x =


Ψ(Φ(x0 )) for some x0 ∈ X, so Ψ(Φ(x))) = Ψ(Φ(Ψ(Φ(x0 )))) = Ψ(Φ(x0 )) = x, so X
is closed.
b) This is just a reformulation of Corollary 2.2.
c) If x ∈ X and y ∈ Y , then Ψ(Φ(x)) = x and Ψ(Φ(y)) = y. 

We speak of the mutually inverse antitone bijections Φ : X → Y and Ψ : Y → X


as the Galois correspondence induced by the Galois connection (Φ, Ψ).

Example: Let K/F be a field extension, and G a subgroup of Aut(K/F ). Then


there is a Galois connection between the set of subextensions of K/F and the set
of subgroups of G, given by

Φ : L → GL = {σ ∈ G | σx = x ∀x ∈ L},

Ψ : H → K H = {x ∈ K | σx = x ∀σ ∈ H}.

Having established the basic results, we will now generally abbreviate the closure
operators Ψ ◦ Φ and Φ ◦ Ψ to x 7→ x and y 7→ y.
3. EXAMPLES OF ANTITONE GALOIS CONNECTIONS 29

2. Lattice properties
Recall that a partially ordered set X is a lattice if for all x1 , x2 ∈ X, there is a
greatest lower bound x1 ∧ x2 and a least upper bound x1 ∨ x2 . A partially ordered
V
set is a complete lattice ifWfor every subset A of X, the greatest lower bound A
and the least upper bound A both exist.
Lemma 2.4. Let (X, Y, Φ, Ψ) be a Galois connection.
a) If X and Y are both lattices, then for all x1 , x2 ∈ X,
Φ(x1 ∧ x2 ) = Φ(x1 ) ∨ Φ(x2 ),
Φ(x2 ∨ x2 ) = Φ(x1 ) ∧ Φ(x2 ).
b) If X and Y are both complete lattices, then for all subsets A ⊂ X,
^ _
Φ( A) = Φ(A),
_ ^
Φ( A) = Φ(A).
Exercise 2.2. Prove Lemma 2.4.
Complete lattices also intervene in this subject in the following way.
Proposition 2.5. Let A be a set and let X = (2A , ⊂) be the power set of A,
partially ordered by inclusion. Let c : X → X be a closure operator. V Then
T the
collection
W c(X)
S of closed subsets of A forms a complete lattice, with S = B∈S B
and S = c( B∈S B).
Exercise 2.3. Prove Proposition 2.5.

3. Examples of Antitone Galois Connections


Example (Indiscretion): Let (X, ≤), (Y, ≤) be partially ordered sets with top el-
ements TX , TY . Define Φ : X → Y , x 7→ TY and Ψ : Y → X, y 7→ TX . Then
(X, Y, Φ, Ψ) is a Galois connection. The induced closure operators are “indiscrete”:
they send every element of X (resp. Y ) to the top element TX (resp. TY ).

Example (Perfection): Let (X, ≤) and (Y, ≤) be anti-isomorphic partially or-


dered sets, i.e., suppose that there exists a bijection Φ : X → Y with x1 ≤ x2 ⇐⇒
Φ(x2 ) ≤ Φ(x1 ). Then the inverse map Ψ : Y → X satisfies y1 ≤ y2 ⇐⇒ Ψ(y2 ) ≤
Ψ(y1 ). Moreover, for x ∈ X, y ∈ Y , x ≤ Ψ(y) ⇐⇒ y = Ψ(Φ(y)) ≤ Φ(x), so
(X, Y, Φ, Ψ) is a Galois connection. Then X = X and Y = Y . As we saw above,
the converse also holds: if X = X and Y = Y then Φ and Ψ are mutually inverse
bijections. Such a Galois connection is called perfect.1

The remaining examples of this section make use of some important ring-theoretic
concepts which will be treated in detail later.

1There is a small paradox here: in purely order-theoretic terms this example is not any more
interesting than the previous one. But in practice given two partially ordered sets it is infinitely
more useful to have a pair of mutually inverse antitone maps running between them than the trivial
operators of the previous example: Galois theory is a shining example! The paradox already shows
up in the distinction between indiscrete spaces and discrete spaces: although neither topology looks
more interesting than the other, the discrete topology is natural and useful (as we shall see...)
whereas the indiscrete topology entirely deserves its alternate name “trivial”.
30 2. GALOIS CONNECTIONS

Example: Let R be a commutative ring. Let X be the set of all ideals of R


and Y = 2Spec R the power set of the set of prime ideals of R. For I ∈ X, put
Φ(I) = V (I) = {p ∈ Spec R | I ⊂ p}.
For V ∈ Y , put \
Ψ(V ) = p.
p∈V
The maps Φ and Ψ are antitone, and for I ∈ X , V ∈ Y,
\
(1) I ⊂ Ψ(V ) ⇐⇒ I ⊂ p ⇐⇒ ∀p ∈ V, I ⊂ p ⇐⇒ V ⊂ Φ(I),
p∈V

so (Φ, Ψ) is a Galois connection. Then X consists of all ideals which can be written
as the intersection of a family of prime ideals. For all I ∈ X,
\
I= p = rad I = {x ∈ R |∃n ∈ Z+ xn ∈ I};
p⊃I

that is, the induced closure operation on X takes any ideal to its radical r(I). In
particular X consists precisely of the radical ideals.
It is not so easy to describe the closure operator on Y or even the subset Y
explicitly, but there is still something nice to say. Since:
(2) V ((0)) = Spec R, V (R) = ∅,

(3) V (I1 ) ∪ V (I2 ) = V (I1 I2 ),


\ X
(4) V (Iα ) = V ( Iα ),
α∈A α∈A

the elements of Y are the closed subsets for a topology, the Zariski topology.

Example: Take R and X as above, but now let S be any set of ideals of R and put
Y = 2S . For I ∈ X, put
Φ(I) = V (I) = {s ∈ S | I ⊂ s}
and for V ∈ Y, put \
Ψ(V ) = s.
s∈V
Once again Φ and Ψ are antitone maps and (1) holds, so we get a Galois connection.
The associated closure operation on X is
\
I 7→ I = s.
s∈S

The relation (4) holds for any S, and the relation (2) holds so long as R ∈ / S. The
verification of (2) for R = Spec R uses the fact that a prime ideal p contains I1 I2
iff it contains I1 or I2 , so as long as S ⊂ Spec S, Y = {V (I) | I ∈ X} are the closed
subsets for a topology on S. This is indeed the topology S inherits as a subspace
of Spec R, so we call it the (relative) Zariski topology.
Various particular choices of S ⊂ Spec R have been considered. Of these the
most important is certainly S = MaxSpec R, the set of all maximal ideals of R.
In this case, X consists of all ideals which can be written as the intersection of
some family of maximal ideals. Such ideals are necessarily radical, but in a general
4. ANTITONE GALOIS CONNECTIONS DECORTICATED: RELATIONS 31

ring not all radical ideals are obtained in this way. Observe that in a general ring
every radical ideal is the intersection of the maximal ideals containing it iff every
prime ideal is the intersection of maximal ideals containing it; a ring satisfying
these equivalent conditions is called a Jacobson ring.

Example: Let k be a field and put R = k[t1 , . . . , tn ]. Then R is a Jacobson ring.


To prove this one needs as prerequisite knowledge Zariski’s Lemma – for every
m ∈ MaxSpec R, the field extension R/m/k is finite – and the proof uses a short
but clever argument: the Rabinowitsch trick.
Suppose that k is algebraically closed. Then Zariski’s Lemma assumes a stronger
form: for all m ∈ MaxSpec R, the k-algebra R/m is equal to k. Let q : R → R/m = k
be the quotient map, and for 1 ≤ i ≤ n, put xi = q(ti ) and x = (x1 , . . . , xn ). It fol-
lows that m contains the ideal mx = ht1 − x1 , . . . , tn − xn i, and since mx is maximal,
m = mx . This gives the following description of the Galois connection between the
set X of ideals of R and Y = 2MaxSpec R , Hilbert’s Nullstellensatz:
(i) Maximal ideals of R are canonically in bijection with n-tuples of points of k,
i.e., with points of affine n-space An/k .
(ii) The closure operation on ideals takes I to its radical ideal rad I.
(iii) The closure operation on subsets of An coincides with topological closure with
respect to the Zariski topology, i.e., the topology on An for which the closed subsets
are the intersections of the zero sets of polynomial functions.

Example: Let K be a field, let X = 2K , let RSpec K be the set of orderings


on K, and let Y = 2RSpec K . Let H : X → Y by
S 7→ H(S) = {P ∈ RSpec K | ∀x ∈ S x >P 0}.
Let Ψ : Y → X by
T 7→ Ψ(T ) = {x ∈ RSpec K |∀P ∈ T x >P 0}.
Then (X, Y, H, Ψ) is a Galois connection.
The set RSpec K carries a natural topology. Namely, we may view any ordering
×
P as an element of {±1}K : P : x ∈ K × 7→ +1 if P (x) > 0 and −1 is P (x) < 0.
×
Giving {±1} the discrete topology and {±1}K , it is a compact (by Tychonoff’s
×
Theorem) zero-dimensional space. It is easy to see that RSpec K embeds in {±1}K
as a closed subspace, and therefore RSpec K is itself compact and zero-dimensional.

Example: Let L be a language, let X be the set of L-theories, and let Y be the
class of all classes C of L-structures, partially ordered by inclusion.2 For a theory
T , let Φ(T ) = CT be the class of all models of T , whereas for a class C, we define
Ψ(C) to be the collection of all sentences ϕ such that for all X ∈ C, X |= ϕ.

4. Antitone Galois Connections Decorticated: Relations


Example: Let S and T be sets, and let R ⊂ S × T be a relation between S and T .
As is traditional, we use the notation xRy for (x, y) ∈ R. For A ⊂ S and y ∈ T ,
we let us write ARy if xRy for all x ∈ A; and dually, for x ∈ S and B ⊂ T , let us
write xRB if xRy for all y ∈ B. Finally, for A ⊂ S, B ⊂ T , let us write ARB if
2Here we are cheating a bit by taking instead of a partially ordered set, a partially ordered
class. We leave it to the interested reader to devise a remedy.
32 2. GALOIS CONNECTIONS

xRy for all x ∈ A and all y ∈ B.


Let X = (2S , ⊂), Y = (2T , ⊂). For A ⊂ S and B ⊂ T , we put
ΦR (A) = {y ∈ T |ARy},
ΨR (B) = {x ∈ S |xRB}.
We claim that GR = (X, Y, ΦR , ΨR ) is a Galois connection. Indeed, it is immediate
that ΦR and ΨR are both antitone maps; moreover, for all A ⊂ S, B ⊂ T we have
A ⊂ ΨR (B) ⇐⇒ ARB ⇐⇒ B ⊂ ΦR (A).
Remarkably, this example includes most of the Galois connections above. Indeed:

• In Example 2.2, take X to be 2K and Y = 2Aut(K/F ) . The induced Galois


connection is the one associated to the relation gx = x on K × Aut(K/F ).
• In Example 2.5, take X to be 2R . The induced Galois connection is the one
associated to the relation x ∈ p on R × Spec R. Similarly for Examples 2.7 and 2.8.
• The Galois connection of Example 2.8 is the one associated to the relation x ∈ P
on K × RSpec K.
• The Galois connection of Example 2.9 is the one associated to the relation X |= ϕ.
Theorem 2.6. Let S and T be sets, let X = (2S , ⊂), Y = (2S , ⊂), and let
G = (X, Y, Φ, Ψ) be any Galois connection. Define a relation R ⊂ S × T by xRy if
y ∈ Φ({x}). Then G = GR .
Proof. Note first that
S X and Y Ware complete lattices, so Lemma 2.4b) applies.
Indeed, for A ⊂ S, A = x∈A {x} = x∈A {x}, so
\ \
Φ(A) = Φ({x}) = {y ∈ T | xRy} = {y ∈ T | ARy} = ΦR (A).
x∈A x∈A

Moreover, since G is a Galois connectionSwe have {x}


W ⊂ Ψ({y}) ⇐⇒ {y} ⊂
Φ({x}) ⇐⇒ xRy. Thus for B ⊂ T , B = y∈B {y} = y∈B {y}, so
\ \
Ψ(B) = Ψ({y}) = {x ∈ S | xRy} = {x ∈ S | xRB} = ΨR (B). 
y∈B y∈A

For any partially ordered set (X, ≤), a downset is a subset Y ⊂ X such that for
all x1 , x2 ∈ X, if x2 ∈ Y and x1 ≤ x2 then x1 ∈ Y . Let D(X) be the collection of
all downsets of X, viewed as a subset of (2X , ⊂). To each x ∈ X we may associate
the principal downset d(x) = {y ∈ X | y ≤ x}. The map d : X → D(X) is an
order embedding; composing this with the inclusion D(X) ⊂ 2X we see that every
partially ordered set embeds into a power set lattice.
Let G = (X, Y, Φ, Ψ) be a Galois connection with X and Y complete lattices.
X Y
Then we may extendV G to a Galois conection between 2 and 2 as follows:
V for A ⊂
X, put Φ(A) = {Φ(x)}x∈A , and simialrly for B ⊂ Y , put Ψ(B) = {Ψ(y)}y∈B .
Thus every Galois connection between complete lattices may be viewed as the Galois
connection induced by a relation between sets.

5. Isotone Galois Connections


Let (X, ≤) and (Y, ≤) be partially ordered sets. An isotone Galois connection
between X and Y is a pair of maps Φ : X → Y and Ψ : Y → X such that:

(IGC1) Φ and Ψ are both isotone maps, and


6. EXAMPLES OF ISOTONE GALOIS CONNECTIONS 33

(IGC2) For all x ∈ X and all y ∈ Y , Φ(x) ≤ y ⇐⇒ x ≤ Ψ(y).

In contrast to the antitone case, this time there is an asymmetry between Φ and
Ψ. We call Φ the lower adjoint and Ψ the upper adjoint.

At the abstract level, the concepts of antitone and isotone Galois connection are
manifestly equivalent.
Exercise 2.4. Let X, Y be partially ordered sets, and let Φ : X → Y , Ψ : Y →
X be functions.
a) Show that (Φ, Ψ) is an antitone Galois connection between X and Y iff (Φ, Ψ)
is an isotone Galois connection between X ∨ and Y .
b) Show that (Φ, Ψ) is an antitone Galois connection between X and Y iff (Ψ, Φ)
is an isotone Galois connection between Y ∨ and X.
If (X, ≤) is a partially ordered set, then a mapping f : X → X is called an interior
operator if it satisfies all of the following properties:

(I1) For all x ∈ X, x ≥ f (x).


(C2) For all x1 , x2 ∈ X, x1 ≤ x2 =⇒ f (x1 ) ≤ f (x2 ).
(C3) For all x ∈ X, f (f (x)) = f (x).
Exercise 2.5. Let (X, ≤) be a partially ordered set, and let f : X → X be a
function. Show: f is a closure operator iff f : X ∨ → X ∨ is an interior operator.
Proposition 2.7. Let (Φ, Ψ) be an isotone Galois connection. Then Ψ ◦ Φ is
an interior operator on (X, ≤), and Φ ◦ Ψ is a closure operator on (Y, ≤).
Proof. By Exercise 2.4, (Φ, Ψ) is an antitone Galois connection between X ∨
and Y , so by Proposition 2.1, Φ ◦ Ψ is a closure operator on Y and Ψ ◦ Φ is a closure
operator on X ∨ and thus, by Exercise 2.5, an interior operator on X. 

6. Examples of Isotone Galois Connections


Example (Galois connection of a function): Let f : S → T be a function. Let
X = (2S , ⊂) and Y = (2T , ⊂). For A ⊂ S and B ⊂ T , put

f∗ (S) = f (S) = {f (s) | s ∈ S}, f ∗ (T ) = f −1 (B) = {s ∈ S | f (s) ∈ B}.


Exercise 2.6. a) Show: (f ∗ , f∗ ) is an isotone Galois connection between 2T
and 2S .
b) Show that the interior operator f∗ ◦ f ∗ : B ⊂ T 7→ B ∩ f (S). In particular the
Galois connection is left perfect iff f is surjective.
c) Show that the Galois connection is right perfect – i.e., f ∗ f∗ A = A for all
A ⊂ S – iff f is injective.
d) Interpret this isotone Galois connection in terms of the “universal” antitone
Galois connection of §2.4.
Example 2.8. (Galois Connection of a Ring Homomorphism): Let f : R → S
be a homomorphism of rings, and let I(R) and I(S) be the lattices of ideals of R
and S. In §1.5 we defined a pushforward map
f∗ : I(R) → I(S), f∗ (I) = hf (I)i
34 2. GALOIS CONNECTIONS

and a pullback map


f ∗ : I(S) → I(R), f ∗ (J) = f −1 (J).
Proposition 2.9. The maps (f ∗ , f∗ ) give an isotone Galois connection between
I(S) and I(T ).
Exercise 2.7. Prove Proposition 2.9.
CHAPTER 3

Modules

1. Basic definitions
Suppose (M, +) is a commutative group. For any m ∈ M and any integer n, one
can make sense of n • m. If n is a positive integer, this means m + · · · + m (n times);
if n = 0 it means 0, and if n is negative, then n • m = −(−n) • m. Thus we have
defined a function • : Z × M → M which enjoys the following properties: for all
n, n1 , n2 ∈ Z, m, m1 , m2 ∈ M , we have

(ZMOD1) 1 • m = m.
(ZMOD2) n • (m1 + m2 ) = n • m1 + n • m2 .
(ZMOD3) (n1 + n2 ) • m = n1 • m + n2 • m.
(ZMOD4) (n1 n2 ) • m = n1 • (n2 • m)

It should be clear that this is some kind of ring-theoretic analogue of a group


action on a set. In fact, consider the slightly more general construction of a monoid
(M, ·) acting on a set S: that is, for all n1 , n2 ∈ M and s ∈ S, we require 1 • s = s
and (n1 n2 ) • s = n1 • (n2 • s).

For a group action G on S, each function g• : S → S is a bijection. For monoidal


actions, this need not hold for all elements: e.g. taking the natural multiplication
action of M = (Z, ·) on S = Z, we find that 0• : Z → {0} is neither injective nor
surjective, ±1• : Z → Z is bijective, and for |n| > 1, n• : Z → Z is injective but not
surjective.

Exercise 3.1. Let • : M × S → S be a monoidal action on a set. Show that for


each unit m ∈ M – i.e., an element for which there exists m0 with mm0 = m0 m = 1
– m• : S → S is a bijection.

Then the above “action” of Z on a commutative group M is in particular a monoidal


action of (Z, ·) on the set M . But it is more: M has an additive structure, and
(ZMOD2) asserts that for each n ∈ Z, n• respects this structure – i.e., is a ho-
momorphism of groups; also (ZMOD3) is a compatibility between the additive
structure on Z and the additive structure on M .

These axioms can be restated in a much more compact form. For a commuta-
tive group M , an endomorphism of M is just a group homomorphism from M
to itself: f : M → M . We write End(M ) for the set of all endomorphisms of
M . But End(M ) has lots of additional structure: for f, g ∈ End(M ) we define
f + g ∈ End(M ) by
(f + g)(m) := f (m) + g(m),
35
36 3. MODULES

i.e., pointwise addition. We can also define f · g ∈ End(M ) by composition:


(f · g)(m) := f (g(m)).
Proposition 3.1. For any commutative group M , the set End(M ) of group
endomorphisms of M , endowed with pointwise addition and multiplication by com-
position, has the structure of a ring.
Exercise 3.2. Prove Proposition 3.1.
Exercise 3.3. Show: End(Z) = Z, and for any n ∈ Z, End(Z/nZ) = Z/nZ.
(More precisely, find canonical isomorphisms.)
These simple examples are potentially misleading: we did not say that the multi-
plication was commutative, and of course there is no reason to expect composition
of functions to be commutative.
Exercise 3.4. a) Show: End(Z/2Z⊕Z/2Z) = M2 (Z/2Z), the (noncom-
mutative!) ring of 2 × 2 matrices with Z/2Z-coefficients.
b) If M is a commutative group and n ∈ Z+ , show End(M n ) = Mn (End(M )).
Now observe that the statement that the action of Z on M satisfes (ZMOD1)
through (ZMOD4) is equivalent to the following much more succinct statement:

For any commutative group M , the map n ∈ Z 7→ (n•) : M → M is a homo-


morphism of rings Z → End(M ).

This generalizes very cleanly: if R is any ring (not necessarily commuative) and
M is a commutative group, a homomorphism • : R → End(M ) will satisfy: for all
r ∈ R, m, m1 , m2 ∈ M :

(LRMOD1) 1 • m = m.
(LRMOD2) r • (m1 + m2 ) = r • m1 + r • m2 .
(LRMOD3) (r1 + r2 ) • m = r1 • m + r2 • m.
(LRMOD4) (r1 r2 ) • m = r1 • (r2 • m).

The terminology here is that such a homomorphism r 7→ (r•) is a left R-module


structure on the commutative group M .

What then is a right R-module structure on M ? The pithy version is that


it is a ring homomorphism from Rop , the opposite ring of R to End(M ). This defi-
nition makes clear (only?) that if R is commutative, there is no difference between
left and right R-module structures. Since our interest is in the commutative case,
we may therefore not worry too much. But for the record:
Exercise 3.5. Show: a homomorphism Rop → End(M ) is equivalent to a
mapping • : M × R → M satisfying
m • 1 = m,
(m1 + m2 ) • r = m1 • r + m2 • r,
m • (r1 + r2 ) = m • r1 + m • r2 ,
m • (r1 r2 ) = (m • r1 ) • r2 .
1. BASIC DEFINITIONS 37

As usual for multiplicative notation, we will generally suppress the bullet, writing
rm for left R-modules and mr for right R-modules.

The calculus of left and right actions is at the same time confusing and some-
what miraculous: it is a somewhat disturbing example of a purely lexicographical
convention that has – or looks like it has – actual mathematical content. Especially,
suppose we have a commutative group M and two rings R and S, such that M si-
multaneously has the structure of a left R-module and a right S-module. Thus we
wish to entertain expressions such as rms for m ∈ M , r ∈ R, s ∈ S. But as stands
this expression is ambiguous: it could mean either
(r • m) • s
or
r • (m • s).
We say that M is an R-S bimodule if both of these expressions agree. Here
is what is strange about this: lexicographically, it is an associativity condition.
But “really” it is a commutativity condition: it expresses the fact that for all
r ∈ R, s ∈ S, (r•) ◦ (•s) = (•s) ◦ (r•): every endomorphism coming from an
element of R commutes with every endomorphism coming from an element of S.
Thus for instance:
Exercise 3.6. Show: any ring R is naturally an R − R-bimodule.
We will not deal with bimodules further in these notes. In fact, when we say R-
module at all, it will be understood to mean a left R-module, and again, since we
shall only be talking about commutative rings soon enough, the distinction between
left and right need not be made at all.

Definition: For M a left R-module, we define its annihilator


ann(M ) = {r ∈ R | ∀m ∈ M, rm = 0}.
Equivalently, ann(M ) is the set of all r such that r· = 0 ∈ End(M ), so that it is
precisely the kernel of the associated ring homomorphism R → End(M ). It follows
that ann(M ) is an ideal of R (note: two-sided, in the noncommutative case).

Definition: A left R-module M is faithful if ann(M ) = 0. Explicitly, this means


that for all 0 6= r ∈ R, there exists m ∈ M such that rm 6= 0.
Exercise 3.7. Let M be an R-module. Show that M has the natural structure
of a faithful R/ ann(M )-module.
Definition: Let M be a left R-module. A submodule of M is a subgroup N
of (M, +) such that RN ⊂ N . The following result is extremely easy and all-
important:
Theorem 3.2. Let R be a ring. The left R-submodules of R are precisely the
left ideals of R.
Exercise 3.8. Prove Theorem 3.2.
Definition: Let M and N be left R-modules. A homomorphism of R-modules is
a homomorphism of commutative groups f : M → N such that for all r ∈ R, m ∈
M, n ∈ N , f (rm) = rf (m).
38 3. MODULES

Exercise 3.9. a) Define an isomorphism of R-modules in the correct way, i.e.,


not as a bijective homomorphism of R-modules.
b) Show that a homomorphism of R-modules is an isomorphism iff it is bijective.
If N is a submodule of a left R-module M , then the quotient group M/N has
a natural R-module structure. More precisely, there is a unique left R-module
structure on M/N such that the quotient map M → M/N is a homomorphism of
R-modules. (Exercise!)
Exercise 3.10. Let I be a two-sided ideal of the not-necessarily-commutative
ring R, so the quotient ring R/I has the structure of a left R-module. Show:
ann(R/I) = I.
In particular, every two-sided ideal of R occurs as the annihilator of a left R-module.

L and {Mi }i∈I a family of R-modules. Con-


Exercise 3.11. a) Let R be a ring
sider the commutative group M = i∈I Mi . Show that putting r(mi ) = (rmi )
makes R into an R-module. Show that the usual inclusion map ιi : Mi → M is a
homomorphism of R-modules.
b) Show that for any R-module N and R-module maps fi : Mi → N , there exists
a unique R-module map f : M → N such that fi = f ◦ ιi for all i ∈ I. Thus M
satisfies the universal mapping property of the direct sum.
Ln
As a matter of notation, for n ∈ Z+ , Rn := i=1 R, R0 = 0.
Exercise 3.12. Work out the analogue of Exercise 3.11 for direct products.
Exercise 3.13. a) Suppose M is an R-module and S is a subset of M . Show
that the intersection of all R-submodules of M containing S is an R-submodule,
and is contained in every R-submodule that contains S. We call it the R-submodule
generated by S and denote it by hSi. If S = {x1 , . . . , xn } is finite, we usually
write hx1 , ..., xn i instead of h{x1 , . . . , xn }i.
P If S = {si }i∈I , show that the R-module generated by S is the set of all sums
b)
i∈J ri si , where J is a finite subset of S.

An R-module M is cyclic (or monogenic) if M = hxi for some x ∈ X.


Exercise 3.14. Show: for an R-module M , the following are equivalent:
(i) M ∼
= R/ ann(M ).
(ii) M is cylic.
Exercise 3.15. Suppose k is a field. Show: the terms “k-module” and “vector
space over k” are synonymous.
One can therefore view the theory of R-modules as a generalization of vector spaces
to arbitrary rings. But really this is something like a zeroth order approximation
of the truth: for a general ring R, the theory of R-modules is incomparably richer
than the theory of vector spaces over a field. There are two explanations for this.
First, even when working with very simple R-modules such as Rn , the usual linear
algebra notions of linear independence, span and basis remain meaningful, but be-
have in unfamiliar ways:

Call a subset S of an R-module M linearly independent if for every finite


subset m1 , . . . , mn of S and any r1 , . . . , rn ∈ R, r1 m1 + . . . + rn mn = 0 implies
1. BASIC DEFINITIONS 39

r1 = . . . = rn = 0. Say that S spans R if the R-submodule generated by S is R,


and finally a basis for an R-module is a subset which is both
L linearly independent
and spanning. For example, for any set I, the R-module i R has a basis ei .

In linear algebra – i.e., when R is a field – every R-module has a basis.1 How-
ever the situation is quite different over a general ring:
Theorem 3.3. a) Let M be an R-module.
L Suppose that S ⊂ R is a basis. Then
M is isomorphic as an R-module to s∈S R. L
be any set, and consider the R-module RS := s∈S R. For each s ∈ S,
b) Let S L
let es ∈ s∈S R be the element whose s-coordinate is 1 and all of whose other
coordinates are 0. Then set {es }s∈S is a basis for RS .
Exercise 3.16. Prove Theorem 3.3.
L
A module which has a basis – so, by the theorem, admits an isomorphism to s∈S R
for some index set S – is called free.
Exercise 3.17. Show: a nonzero free R-module is faithful.
Let us examine the case of modules over R = Z, i.e., of commutative groups. Here
the term free commutative group is synonymous with “free Z-module”. Needless
to say (right?), not all commutative groups are free: for any integer n > 1, Z/nZ
is not free, since it has nonzero annihilator nZ. Thus Z/nZ does not have a basis
as a Z-module, and indeed has no nonempty linearly independent subsets!
Proposition 3.4. For a commutative ring R, the following are equivalent:
(i) Every R-module is free.
(ii) R is a field.
Proof. As discussed above, (ii) =⇒ (i) is a fundamental theorem of linear
algebra, so we need only concern ourselves with the converse. But if R is not a field,
then there exists a nonzero proper ideal I, and then R/I is a nontrivial R-module
with 0 6= I = ann(R/I), so by Exercise 3.16 R/I is not free. 
Remark: If R is a not-necessarily-commutative ring such that every left R-module
is free, then the above argument shows R has no nonzero proper twosided ideals,
so is what is called a simple ring. But a noncommutative simple ring may still
admit a nonfree module. For instance, let k be a field and take R = M2 (k), the
2 × 2 matrix ring over k. Then k ⊕ k is a left R-module which is not free. How-
ever, suppose R is a ring with no proper nontrivial one-sided ideals. Then R is a
division ring – i.e., every nonzero element of R is a unit – and every R-module is free.

In linear algebra – i.e., when R is a field – every linearly independent subset of


an R-module can be extended to a basis. Over a general ring this does not hold
even for free R-modules. For instance, take R = M = Z. A moment’s thought
reveals that the only two bases are {1} and {−1}, whereas the linearly independent
sets are precisely the singleton sets {n} as n ranges over the nonzero integers.

Note well the form of Proposition 3.4: we assume that R is a commutative ring for
1This uses, and is in fact equivalent to, the Axiom of Choice, but the special case that any
vector space with a finite spanning set has a basis does not.
40 3. MODULES

which R-modules satisfy some nice property, and we deduce a result on the struc-
ture of R. Such “inverse problems” have a broad appeal throughout mathematics
and provide one of the major motivations for studying modules above and beyond
their linear algebraic origins. We will see other such characterizations later on.

2. Finitely presented modules


One of the major differences between commutative groups and noncommutative
groups is that a subgroup N of a finitely generated commutative group M remains
finitely generated, and indeed, the minimal number of generators of the subgroup
N cannot exceed the minimal number of generators of M , whereas this is not true
for nonabelian groups: e.g. the free group of rank 2 has as subgroups free groups
of every rank 0 ≤ r ≤ ℵ0 . (For instance, the commutator subgroup is not finitely
generated.)

Since an commutative group is a Z-module and every R-module has an underlying


commutative group structure, one might well expect the situation for R-modules to
be similar to that of commutative groups. We will see later that this is true in many
but not all cases: an R-module is called Noetherian if all of its submodules are
finitely generated. Certainly a Noetherian module is itself finitely generated. The
basic fact here – which we will prove in §8.7 – is a partial converse: if the ring R
is Noetherian, any finitely generated R-module is Noetherian. We can already see
that the Noetherianity of R is necessary: if R is not Noetherian, then by definition
there exists an ideal I of R which is not finitely generated, and this is nothing else
than a non-finitely generated R-submodule of R (which is itself generated by the
single element 1.) Thus the aforementioned fact about subgroups of finitely gener-
ated commutative groups being finitely generated holds because Z is a Noetherian
ring.

When R is not Noetherian, it becomes necessary to impose stronger conditions


than finite generation on modules. One such condition indeed comes from group
theory: recall that a group G is finitely presented if it is isomorphic to the quo-
tient of a finitely generated free group F by the least normal subgroup N generated
by a finite subset x1 , . . . , xm of F .
Proposition 3.5. For a finitely generated R-module M , the following are
equivalent:
(i) There exist non-negative integers m, n and an exact sequence
Rm → Rn → M → 0.
(ii) M is the quotient of a finitely generated free R-module by a finitely generated
submodule.
A module M satisfying these equivalent conditions is said to be finitely presented.
Proof. (i) =⇒ (ii) is immediate. Conversely, let M = Rn /N where N is
finitely generated. Then there exists a surjection Rm → N and thus the sequence
Rm → Rn → M → 0
is exact. 
2. FINITELY PRESENTED MODULES 41

Proposition 3.6. Let


ψ φ
0→K→N →M →0
be a short exact sequence of R-modules, with M finitely presented and N finitely
generated. Then K is finitely generated.
Proof. (Matsumura) By definition of finitely presented, we can place M in
an exact sequence
f
(5) Rm → Rn → M → 0
for some m, n ∈ N. For 1 ≤ i ≤ n, let ei be the ith standard basis element of M ,
let mi = f (ei ) be the image in M , and choose ni ∈ N any element in φ−1 (mi ).
Then there is a unique R-module homomorphism α : Rn → N given by α(ei ) = ni ,
which restricts to an R-module homomorphism β : B m → K. Altogether we get a
commutative diagram
f
Rm −→ Rn −→ M −→ 0
ψ φ
0 −→ K −→ N −→ M.
The rest of the proof is essentially a diagram chase. Suppose N = hξ1 , . . . , ξk iR ,
and choose v1 , . . . , vk ∈ Rn such that φ(ξi ) = f (vi ). Put
ξi0 = ξi − α(vi ).
Then ϕ(ξi0 ) = 0, so there exist unique ηi ∈ K such that
ξi0 = ψ(ηi ).
We claim that K is generated as an R-module by β(Rm ) and η1 , . . . , ηk and thus
is finitely generated. Indeed, for η ∈ K, there are r1 , . . . , rk ∈ R such that
X
ψ(η) = ri ξi .
i

Then X X X
ψ(η − ri ηi ) = ri (ξi − ξi0 ) = α( ri vi ).
i i i
Since X X
0 = φ(α( ri vi )) = f ( ri vi ),
i i
m
P
we may write = g(u) with u ∈ R . Then
i ri vi
X X
ψ(β(u)) = α(g(u)) = α( ri vi ) = ψ(η − ri ηi ).
i i

Since ψ is injective, we conclude


X
η = β(u) + ri ηi . 
i

Exercise 3.18. Let 0 → M 0 → M → M 00 → 0 be a short exact sequence of


R-modules.
a) Show: if M 0 and M 00 are both finitely presented, so is M .
b) Show”: if M is finitely presented and M 0 is finitely generated, then M 00 is finitely
presented.
42 3. MODULES

A stronger condition still is the following: an R-module M is coherent if it is


finitely generated and every finitely generated submodule is finitely presented. Ev-
idently coherent implies finitely presented implies finitely generated, and all three
coincide over a Noetherian ring. The significance of coherence lies in the following:
Theorem 3.7. Let R be a not-necessarily-commutative ring.
a) The category of all left R-modules is an abelian category.
b) The category of all coherent left R-modules is an abelian category.
c) In particular, if R is left Noetherian, the category of all finitely generated left
R-modules is an abelian category.
d) There exists a commutative ring R for which the category of all finitely generated
(left) R-modules is not abelian.
The proof of this result – and even an explanation of the term “abelian category” is
beyond the scope of this text, so this is definitely an ancillary remark. Nevertheless
we hope that it will be of some use to students of algebraic geometry: for instance,
it explains the fact that Hartshorne only defines coherent sheaves of modules on a
scheme X in the case that the scheme is Noetherian and suggests the correct (and
more subtle) definition in the non-Noetherian case.

3. Torsion and torsionfree modules


Let R be a domain, and let M be an R-module. An element x ∈ M is said to be
torsion if there exists 0 6= a ∈ R such that ax = 0. Equivalently, the annihilator
ann(x) = {a ∈ R | ax = 0} is a nonzero ideal of R. We define M [tors] to be the set
of all torsion elements of M . It is immediate to see that M [tors] is a submodule
of M . We say that M is a torsion R-module if M = M [tors] and that M is
torsionfree if M [tors] = 0.
Exercise 3.19. Let 0 → M1 → M → M2 → 0 be an exact sequence.
a) Show that if M is torsion, so are M1 and M2 .
b) If M1 and M2 are torsion modules, must M be torsion?
c) Show that if M is torsionfree, show that so is M1 , but M2 need not be.
d) If M1 and M2 are torsionfree, must M be torsionfree?
Proposition 3.8. Let R be a domain and M an R-module.
a) The quotient M/M [tors] is torsionfree.
b) If M is finitely generated, the following are equivalent:
(i) M embeds in a finitely generated free R-module.
(ii) M is torsionfree.
Proof. a) Put N = M/M [tors], and let x ∈ N be such that there exists
0 6= a ∈ R with ax = 0. Let x be any lift of x to M ; then there exists t ∈ M [tors]
such that ax = t. By definition of torsion, there exists a0 ∈ R such that a0 t = 0, so
a0 ax = a0 t = 0. Since R is a domain, a0 a is nonzero, so x ∈ M [tors] and x = 0.
b) (i) =⇒ (ii) is very easy: free modules are torsionfree and submodules of
torsionfree modules are torsionfree.
(ii) =⇒ (i): We may assume M 6= 0. Let M = hx1 , . . . , xr i with r ≥ 1 and all the
xi are nonzero. Further, after reordering the xi ’s if necessary, there exists a unique
s, 1 ≤ s ≤ r, such that {x1 , . . . , xs } is linearly independent over R but for all i with
s < i ≤ r, {x1 , . . . xs , xi } is linearly dependent over R. Then F = hx1 , . . . , xs i ∼
= Rs ,
so we are done if s = r. If s < r, then for each i > s there exists 0 6= ai ∈ R such
4. TENSOR AND HOM 43

Q
that ai xi ∈ F . Put a = s<i≤r ai : then aM ⊂ F . Let [a] : M → M denote
multiplication by a. Since M is torsionfree, [a] is injective hence gives an R-module
isomorphism from M to a submodule of the finitely generated free module F . 
Exercise 3.20. Show: the torsionfree Z-module (Q, +) is not isomorphic to a
submodule of any finitely generated free Z-module.
(Thus – even for very nice rings! – the hypothesis of finite generation is necessary
in Proposition 3.8.)

4. Tensor and Hom


4.1. Tensor products.

We assume that the reader has some prior familiarity with tensor products, say
of vector spaces and/or of abelian groups. The first is an instance of tensor prod-
ucts of k-modules, for some field k, and the second is an instance of tensor products
of Z-modules. We want to give a general definition of M ⊗R N , where M and N
are two R-modules.

There are two ways to view the tensor product construction: as a solution to a
universal mapping problem, and as a generators and relations construction. They
are quite complementary, so it is a matter of taste as to which one takes as “the”
definition. So we will follow our taste by introducing the mapping problem first:

Suppose M , N , P are R-modules. By an R-bilinear map f : M × N → P we


mean a function which is separately R-linear in each variable: for all m ∈ M , the
mapping n 7→ f (m, n) is R-linear, and for each n ∈ N , the mapping m 7→ f (m, n) is
R-linear. Now consider all pairs (T, ι), where T is an R-module and ι : M × N → T
is an R-bilinear map. A morphism from (T, ι) to (T 0 , ι0 ) will be an R-module homo-
morphism h : T → T 0 such that ι0 = h ◦ ι. By definition, a tensor product M ⊗R N
is an initial object in this category: i.e., it comes equipped with an R-bilinear map
M × N raM ⊗R N such that any R-bilinear map f : M × N → P factors through
it. As usual, the initial object of a category is unique up to unique isomorphism
provided it exists.
As for the existence, we fall back on the generators and relations construction.
Namely, we begin with the free R-module F whose basis is M × N , and we write
the basis elements (purely formally) as m ⊗ n. We then take the quotient by the
submodule generated by the following relations R:
(x + x0 ) ⊗ y − x ⊗ y − x0 ⊗ y,
x ⊗ (y + y 0 ) − x ⊗ y − x ⊗ y 0 ,
(ax) ⊗ y − a(x ⊗ y),
x ⊗ (ay) − a(x ⊗ y).
It is then easy to see that the quotient map M × N → F/N satisfies all the prop-
erties of a tensor product (details left to the reader).

Note that the general element of M ⊗R N is not a single element of the form x ⊗ y
but rather a finite sum of such elements. (Indeed, from the free R-module, every
element can be represented by a finite R-linear combination of elements of the form
44 3. MODULES

x ⊗ y, but the last two defining relations in the tensor product allow us to change
ri (x⊗y) to either (ri x)⊗y or x⊗(ri y).) Of course, this representation of an element
of the tensor product need not be (and will never be, except in trivial cases) unique.

One can also take the tensor product of R-algebras: if R is a (commutative!) ring
and A and B are commutative R-algebras, then on the tensor product A ⊗R B we
have a naturally defined product, induced by (a1 ⊗ b1 ) · (a2 ⊗ b2 ) := (a1 a2 ⊗ b1 b2 ).
We have to check that this is well-defined, a task which we leave to the reader (or
see [AM, pp. 30-31])). The tensor product of algebras is a powerful tool – e.g.
in the structure theory of finite-dimensional algebras over a field, or in the theory
of linear disjointness of field extensions – and is given misleadingly short shrift in
most elementary treatments.

Base change: Suppose that M is an R-module and f : R → S is a ring homo-


morphism. Then S is in particular an R-module, so that we can form the tensor
product S P⊗R M . This is still
P an R-module, but it is also an S-module in an evident
way: s • ( i si ⊗ mi ) := i ssi ⊗ mi . This is process is variously called scalar
extension, base extension or base change. Note that this process is functo-
rial, in the following sense: if f : M → M 0 is an R-algebra homomorphism, then
there exists an induced S-algebra homomorphism S ⊗R M → S ⊗R M 0 , given by
s ⊗ m 7→ s ⊗ f (m).
Exercise 3.21. If M is a finitely generated R-module and f : R → S is a ring
homomorphism, then S ⊗R M is a finitely generated S-module.
Exercise 3.22. Let A and B be rings, M an A-module, P a B-module, and N
an (A, B)-bimodule. Then M ⊗A N is naturally a B-module, N ⊗B P is naturally
an A-module, and
(M ⊗A N ) ⊗B P ∼
= M ⊗A (N ⊗B P ).
Exercise 3.23. Let R be a commutative ring, I an ideal of R and M an R-
module.
a) Show: there is a well-defined R-bilinear map R/I × M → M/IM given by
(r + I, m) 7→ rm + I. Thus there is an induced homomorphism of R-modules
ϕ : R/I ⊗R M → M/IM.
b) Show: ϕ is an isomorphism of R-modules.
Proposition 3.9. Let R be a commutative ring, M an R-module and {Ni }i∈I
a directed system of R-modules. Then the R-modules lim(M ⊗Ni ) and M ⊗(lim Ni )
−→ −→
are canonically isomorphic.
Exercise 3.24. Prove Proposition 3.22.
Exercise 3.25. Let M and N be R-modules.
a) Show ann M ⊗ N ⊃ ann M + ann N .
b) Suppose M and N are cyclic R-modules. Show: M ⊗ N is cyclic and
ann(M ⊗ N ) = ann M + ann N . Equivalently, show that for all ideals I, J
of R we have
R/I ⊗ R/J = R/(I + J).
5. PROJECTIVE MODULES 45

c) Deduce: for all m, n ∈ Z we have


Z/mZ ⊗ Z/nZ = Z/hm, niZ.
d) Let2 k be a field, let R := k[t2 , t3 ], M := k[t], I = t2 R, N := R/I. Show:
t3 ∈ ann(M/IM ) \ ann M + I = I
and thus
ann(M ⊗ N ) ) ann M + ann N.

5. Projective modules
5.1. Basic equivalences.
Proposition 3.10. For an R-module P , the following are equivalent:
(i) There exists an R-module Q such that P ⊕ Q is a free R-module.
(ii) If π : M → N is a surjective R-module homomorphism and ϕ : P → N is a
homomorphism, then there exists at least one R-module homomorphism Φ : P → M
such that ϕ = π ◦ Φ.
(iii) If π : M → N is a surjection, then the natural map Hom(P, M ) → Hom(P, N )
given by Φ 7→ π ◦ Φ is surjective.
(iv) The functor Hom(P, ) is exact.
(v) Any short exact sequence of R-modules
q
0→N →M →P →0
splits: there exists an R-module map σ : P → M such that q ◦ σ = 1P and thus an
internal direct sum decomposition M = N ⊕ σ(P ).
A module satisfying these equivalent conditions is called projective.
Proof. (i) =⇒ (ii): Let F ∼ = P ⊕ Q be a free module. Let {fi } be a free basis
for F and let {pi } be the corresponding generating set for P , where pi is the image
of fi under the natural projection P ⊕Q → P . Put ni = ϕ(pi ). By surjectivity of π,
let mi ∈ π −1 (ni ). By the freeness of F , there is a unique R-module homomorphism
h : F → M carrying each fi to mi . Pull h back to P via the natural inclusion
P ,→ F . Then h : P → M is such that π ◦ f = ϕ.
(ii) =⇒ (i): As for any R-module, there exists a free R-module F and a surjection
π : F → P . Applying (ii) with N = p and ϕ : P → N the identity map, we get a
homomorphism Φ : P → F such that π ◦ ϕ = 1P . It follows that F = Φ(P ) ⊕ ker(π)
is an internal direct sum decomposition.
(ii) ⇐⇒ (iii): (iii) is nothing more than a restatement of (ii), as we leave it to the
reader to check.
(iii) ⇐⇒ (iv): To spell out (iv), it says: if
0 → M 0 → M → M 00 → 0
is a short exact sequence of R-modules, then the corresponding sequence
0 → Hom(P, M 0 ) → Hom(P, M ) → Hom(P, M 00 ) → 0
is exact. Now for any R-module P , the sequence
0 → Hom(P, M 0 ) → Hom(P, M ) → Hom(P, M 00 )

2(This is taken from https://math.stackexchange.com/questions/79538/.


46 3. MODULES

is exact – i.e., Hom(P, ) is left exact – so (iv) amounts to: for any surjection
M → M 00 , the corresponding map Hom(P, M ) → Hom(P, M 00 ) is surjective, and
this is condition (iii).
(ii) =⇒ (v): Given
q
0 → N → M → P → 0,
we apply (ii) to the identity map 1P : P → P and the surjection q : M → P ,
getting a map σ : P → M such that q ◦ σ = 1P , so σ is a section as required.
(v) =⇒ (i): Choosing a set of generators for P gives rise to a surjective ho-
momorphism q : F → P from a free R-module F to P and thus a short exact
sequence
q
0 → Ker q → F → P → 0.
By hypothesis, there exists a section σ : P → F and thus an internal direct sum
decomposition F ∼ = Ker(q) ⊕ σ(P ) ∼
= Ker(q) ⊕ P . 
Exercise 3.26. Give a direct proof that (v) =⇒ (ii) in Proposition 3.10.
(Suggestion: Given the surjection q : M → N and the map π : P → N , form the
short exact sequence 0 → K → M → N → 0 and show that it is mapped to by a
short exact sequence 0 → K → M ×N P → P → 0, where
M ×N P = {(x, y) ∈ M × P | q(x) = π(y)}
is the fiber product of M and P over N .)
Exercise 3.27. Use Proposition 3.10 to show, several times over, that a free
R-module is projective.

L Let {Mi }i∈I be an index family of R-modules. Show that the


Exercise 3.28.
direct sum M = i∈I Mi is projective iff each Mi is projective.
Exercise 3.29. a) Show: the tensor product of two free R-modules is free.
b) Show: the tensor product of two projective R-modules is projective.
Exercise 3.30. Show: a finitely generated projective module is finitely pre-
sented. (Hint: the problem is that over a not-necessarily-Noetherian ring, a sub-
module of a finitely generated module need not be finitely generated. However, a
direct summand of a finitely generated module is always finitely generated: why?)
5.2. Linear algebraic characterization of projective modules.

Let R be a commutative ring, n ∈ Z+ , and let P be an element of the (non-


commutative!) ring Mn (R) of n × n matrices with entries in R such that P 2 = P .
There are several names for such a matrix. The pure algebraist would call such a
matrix idempotent, for that is the name of an element in any ring which is equal
to its square. A geometrically minded algebraist however may call such a matrix
a projection, the idea being that the corresponding R-module endomorphism of
Rn “projects” Rn onto the submodule P (Rn ).
Proposition 3.11. An R-module M is finitely generated and projective iff it
is, up to isomorphism, the image of a projection: i.e., iff there exists n ∈ Z+ and
a matrix P ∈ Mn (R) with P = P 2 such that M ∼ = P (Rn ).
Proof. Suppose first that M is a finitely generated projective R-module. Since
M is finitely generated,, there exists n ∈ Z+ and a surjective R-module homorphism
5. PROJECTIVE MODULES 47

π : Rn → M . Since M is projective, this homomorphism has a section σ : M → Rn ,


and we may thus write Rn = σ(M ) ⊕ M 0 . Put P = σ ◦ π ∈ EndR (Rn ). Then
P (Rn ) = σ(π(Rn )) = σ(M ) ∼
= M and
P 2 = σ ◦ (π ◦ σ) ◦ π = σ ◦ 1M ◦ π = σ ◦ π = P.
Conversely, suppose that there exists P ∈ EndR (Rn ) with P 2 = P and let M ∼ =
P (Rn ). Then – since P (1 − P ) = 0 – Rn = P (Rn ) ⊕ (1 − P )(Rn ), exhibiting P (Rn )
as a direct summand of a free module.3 
5.3. The Dual Basis Lemma.
Proposition 3.12. (Dual Basis Lemma) For an R-module M , the following
are equivalent:
(i) There exists an index set I, elements {ai }i∈I of M and homomorphisms {fi :
M → R}i∈I such that for each a ∈ M , {i ∈ I | fi (a) 6= 0} is finite, and
X
a= fi (a)ai .
i∈I

(ii) M is projective.
Proof. (i) =⇒ (ii): Let F be the free R-module with basis elements {ei }i∈I ,
P f : F → M by f (ei ) = ai . Then the map ι : M → F given by
and define
ι(a) = i∈I fi (a)ei is a section
L of f , so M is a direct summand of F .
(ii) =⇒ (i): Let f : F = i∈I R → M be an epimorphism from a free R-module
onto M . Since M is projective, there exists a section ι : M ,→ F . If {ei }i∈I is the
standard basis of F , then for all a ∈ M , the expression
X
ι(a) = fi (a)ei
i∈I
defines the necessary family of functions fi : M → R. 
Exercise 3.31. Let P be a projective R-module. Show: one can can find a
finite index set I satisfying condition (i) of the Dual Basis Lemma iff P is finitely
generated.
5.4. Projective versus free.

Having established some basic facts about projective modules, we should now seek
examples in nature: which modules are projective? By Exercise 3.25 any free mod-
ule is projective. But this surely counts as a not very interesting example! Indeed
the following turns out to be one of the deepest questions of the subject.
Question 1. When is a projective module free?
We want to give examples to show that the answer to Question 1 is not “always”.
But even by giving examples one wades into somewhat deep waters. The following
is the one truly “easy” example of a non-free projective module I know.

Example: Suppose R1 and R2 are nontrivial rings. Then the product R = R1 × R2


admits nonfree projective modules. Indeed, let P be the ideal R1 × {0} and Q the
ideal {0} × R2 . Since R = P ⊕ Q, P and Q are projective. On the other hand P
3This part of the proof redeems the pure algebraist: this the decomposition afforded by the
pair of orthogonal idempotents P, 1 − P .
48 3. MODULES

cannot be free because taking e := (0, 1) ∈ R, we have eP = 0, whereas eF 6= 0 for


any nonzero free R-module F (and of course, Q is not free either for similar reasons).

Question 1 may be construed in various ways. One way is to ask for the class
of rings over which every projective module is free, or over which every finitely gen-
erated projective module is free. I actually do not myself know a complete answer
to this question, but there are many interesting and important special cases.

Recall the following result from undergraduate algebra.


Theorem 3.13. A finitely generated module over a PID is free iff it is torsion-
free.
Of course submodules of torsionfree modules are torsionfree, so projective implies
torsionfree. We deduce:
Corollary 3.14. A finitely generated projective module over a PID is free.
Theorem 3.13 does not extend to all torsionfree modules: for instance, the Z-module
Q is torsionfree but not free. However Corollary 3.14 does extend to all modules
over a PID. The proof requires transfinite methods and is given in §3.10.

Recall that a ring R is local if it has a unique maximal ideal. It is convenient


to reserve the notation m for the unique maximal ideal of a local ring and speak of
“the local ring (R, m)”. We want to show that every finitely generated projective
module over a local ring is free. First a few preliminaries.

Let f : R → S be a homomorphism of rings. Then necessarily f induces a ho-


momorphism f × : R× → S × on unit groups: if xy = 1, then f (x)f (y) = f (1) = 1,
so units get mapped to units. But what about the converse: if x ∈ R is such that
f (x) is a unit in S, must x be a unit in R?
It’s a nice idea, but it’s easy to see that this need not be the case. For instance,
let a > 1 be any positive integer. Then a is not a unit of Z, but for each prime
p > a, the image of a in the quotient ring Z/pZ is a unit. Too bad! Let us not give
up so soon: a conjecture may fail, but a definition cannot: say a homomorphism
f : R → S of rings is unit-faithful if for all x ∈ R, f (x) ∈ S × =⇒ x ∈ R× .
Lemma 3.15. If (R, m) is a local ring, the quotient map q : R → R/m is
unit-faithful.
Proof. An element of any ring is a unit iff it is contained in no maximal ideal,
so in a local ring we have R× = R \ m. Moreover, since m is maximal, R/m is a
field. Thus, for x ∈ R,
q(x) ∈ (R/m)× ⇐⇒ x ∈
/ m ⇐⇒ x ∈ R× . 
Later we will see a generalization: if J is any ideal contained the Jacobson radical
of R, then q : R → R/J is unit-faithful.
Theorem 3.16. A finitely generated projective module over a local ring is free.
Proof. Let P be a finitely generated projective module over the local ring
(R, m). We may find Q and n ∈ Z+ such that P ⊕ Q = Rn . Now tensor with
R/m: we get a direct sum decomposition P/mP ⊕ Q/mQ = (R/m)n . Since R/m
5. PROJECTIVE MODULES 49

is a field, all R/m-modules are free. Choose bases {pi } for P/mP and {qj } for
Q/mQ, and for all i, j, lift each pi to an element pi of P and each qj to an element
qj of Q. Consider the n × n matrix A with coefficients in R whose columns are
p1 , . . . , pa , q1 , . . . , qb . The reduction modulo m of A is a matrix over the field R/m
whose columns form a basis for (R/m)n , so its determinant is a unit in (R/m)× .
Since det(M (mod m)) = det(M ) (mod m), Lemma 3.15 implies that det(M ) ∈
R× , i.e., M is invertible. But this means that its columns are linearly independent.
By a consequence of Nakayama’s Lemma (Corollary 3.42) we have that p1 , . . . , pa
spans P , so in fact it forms a basis for P . 
Once again, in Section 3.9 this result will be improved upon: it is a celebrated
theorem of Kaplansky that any projective module over a local ring is free.
Much more interesting is an example of a finitely generated projective, nonfree
module over a domain. Probably the first such examples come from nonprincipal
ideals in rings of integers of number fields with class number greater than 1. To
give such an example with proof of its projectivity this early in the day, we require
a little preparation.4

Two ideals I and J in a ring R are comaximal if I + J = R. More generally,


a family {Ii } of ideals in a ring is pairwise comaximal if for all i 6= j, I + J = R.
Lemma 3.17. Let I, J, K1 , . . . , Kn be ideals in the ring R.
a) We have (I + J)(I ∩ J) ⊂ IJ.
b) If I and J are comaximal, IJ = I ∩ J.
c) If I + Ki = R for all 1 ≤ i ≤ n, then I + K1 · · · Kn = R.
Proof. a) (I + J)(I ∩ J) = I(I ∩ J) + J(I ∩ J) ⊂ IJ + IJ = IJ.
b) If I + J = R, the identity of part a) becomes I ∩ J ⊂ IJ. Since the converse
inclusion is valid for all I and J, the conclusion follows. c) We go by induction on
n, the case n = 1 being trivial. If n = 2, then for i = 1, 2, let ai ∈ I and bi ∈ Ki be
such that 1 = ai + bi . Then
1 = a1 + a2 − a1 a2 + b1 b2 ∈ I + K1 K2 .
Now assume n ≥ 3 and that the result holds for n − 1. By induction,
I + K1 · · · Kn−1 = R.
and by hypothesis I + Kn = R, so by the n = 2 case we have
I + K1 · · · Kn = R. 
Proposition 3.18. Let I and J be comaximal ideals in a domain R, and
consider the R-module map q : I ⊕ J → R given by (x, y) 7→ x + y. Then:
a) The map q is surjective.
b) Ker(q) = {(x, −x) | x ∈ I ∩ J}, hence is isomorphic as an R-module to I ∩ J.
c) We have an isomorphism of R-modules
I ⊕J =∼ IJ ⊕ R.
d) Thus if IJ is a principal ideal, I and J are projective modules.
4Here we wish to acknowledge our indebtedness to K. Conrad: we took our
inspiration for Proposition 3.18 and the following Exercise from Example 3.1 of
http://www.math.uconn.edu/∼kconrad/blurbs/linmultialg/splittingmodules.pdf.
50 3. MODULES

Proof. It is clear that for any ideals I and J, the image of the map q is the
ideal I + J, and we are assuming I + J = R, whence part a).
Part b) is essentially immediate: details are left to the reader.
Combining parts a) and b) we get a short exact sequence
0 → I ∩ J → I ⊕ J → R → 0.
But R is free, hence projective, and thus the sequence splits, giving part c). Finally,
a nonzero principal ideal (x) in a domain R is isomorphic as an R-module to R
itself: indeed, multiplication by x gives the isomorphism R → (x). So if IJ is
principal, I ⊕ J ∼
= R2 and I and J are both direct summands of a free module. 
In particular, if we can find in a domain R two comaximal nonprincipal ideals I
and J with IJ principal, then I and J are finitely generated projective nonfree
R-modules. The following exercise asks you to work through an explicit example.

Exercise 3.32. Let R = Z[ −5], and put
√ √
p1 = h3, 1 + −5i, p2 = h3, 1 − −5i.
a) Show that R/p1 ∼ = R/p2 ∼ = Z/3Z, so p1 and p2 are maximal ideals of R.
b) Show that p1 + p2 = R (or equivalently, that p1 6= p2 ).
c) Show that p1 p2 = (3).
d) Show that neither p1 nor p2 is principal.
√ √
(Suggestion: show that if p1 = (x + −5y) then p2 = (x − −5y) and thus there
are integers x, y such that x2 + 5y 2 = ±3.)
e) Conclude that p1 and p2 are (in fact isomorphic) nonfree finitely generated pro-
jective modules over the domain R.
f ) Show that p2 is principal, and thus that the class of p in K
^0 (R) is 2-torsion.

This construction looks very specific, and the number-theoretically inclined reader
is warmly invited to play around with other quadratic rings and more general rings
of integers of number fields to try to figure out what is really going on. From our
perspective, we will (much later on) gain a deeper understanding of this in terms
of the concepts of invertible ideals, the Picard group and Dedekind domains.

Example: Let X be a compact space, and let C(X) be the ring of continuous
real-valued functions on X. The basic structure of these rings is studied in §5.2.
Let E → X be a real topological vector bundle over X. Then the group Γ(E)
of global sections is naturally a module over C(X). In fact it is a finitely gener-
ated projective module, and all finitely generated projective C(X)-modules arise
faithfully in this way: the global section functor gives a categorical equivalence
between vector bundles on X and finitely generated projective modules over C(X).
This is a celebrated theorem of R.G. Swan, and Section X is devoted to giving
a self-contained discussion of it, starting from the definition of a vector bundle.
In particular, via Swan’s Theorem basic results on the tangent bundles of com-
pact manifolds translate into examples of finitely generated projective modules: for
instance, an Euler characteristic argument shows that the tangent bundle of any
even-dimensional sphere S 2k is nontrivial, and thus Γ(T S 2k ) is a finitely generated
nonfree C(S 2k )-module! Following Swan, we will show that examples of nonfree
projective modules over more traditional rings like finitely generated R-algebras
follow from examples like these.
5. PROJECTIVE MODULES 51

Example: Let k be a field and R = k[t1 , . . . , tn ] be the polynomial ring over k


in n indeterminates. When n = 1, R is a PID, so indeed every finitely generated
R-module is projective. For n > 1, the situation is much less clear, but the problem
of freeness of finitely generated projective R-modules can be stated geometrically as
follows: is any algebraic vector bundle on affine n-space An/k algebraically trivial?
When k = C, the space An/C = Cn in its usual, Euclidean topology is contractible,
which by basic topology implies that any continuous C-vector bundle on An is (con-
tinuously) trivial. Moreover, relatively classical complex variable theory shows that
any holomorphic vector bundle on An is (holomorphically) trivial. But asking the
transition functions and the trivialization to be algebraic – i.e., polynomial func-
tions – is a much more stringent problem. In his landmark 1955 paper FAC, J.-P.
Serre noted that this natural problem remained open for algebraic vector bundles:
he was able to prove only the weaker result that a finitely generated projective
R-module M is stably free – i.e., there exists a finitely generated free module
M such that M ⊕ F is free. This became known as Serre’s Conjecture (to
his dismay) and was finally resolved independently in 1976 by D. Quillen [Qu76]
and A. Suslin [Su76]: indeed, every finitely generated projective R-module is free.
Quillen received the Fields Medal in 1978. Fields Medals are not awarded for the
solution of any single problem, but the prize committee writes an official docu-
ment describing the work of each winner that they found particularly meritorious.
In this case, it was made clear that Quillen’s resolution of Serre’s Conjecture was
one of the reasons he received the prize. All this for modules over a polynomial ring!

For more information on Serre’s Conjecture, the reader can do no better than
to consult a recent book of T.Y. Lam [La06].
Exercise 3.33. (K0 (R)): From a commutative ring R, we will construct an-
other commutative ring K0 (R) whose elements correspond to formal differences of
finite rank projective modules. More precisely:
a) Let M0 (R) denote the set of all isomorphism classes of finitely generated projec-
tive modules. For finitely generated projective modules P and Q we define
[P ] + [Q] = [P ⊕ Q],
[P ] · [Q] = [P ⊗ Q].
Check that this construction is well-defined on isomorphism classes and endows
M0 (R) with the structure of a commutative semiring with unity. What are the
additive and mulitplicative identity elements?
b) Define K0 (R) as the Grothendieck group of M0 (R), i.e., as the group completion
of the commutative monoid M0 (R). Convince yourself that K0 (R) has the structure
of a semiring. The elements are of the form [P ] − [Q], and we have [P1 ] − [Q1 ] =
[P2 ] − [Q2 ] ⇐⇒ there exists a finitely generated projective R-module M with
P1 ⊕ Q2 ⊕ M ∼ = P2 ⊕ Q1 ⊕ M.
In particular, if P and Q are projective modules, then [P ] = [Q] in K0 (R) iff [P ]
and [Q] are stably isomorphic, i.e., iff they become isomorphic after taking the
direct sum with some other finitely generated projective module M . c) Show: we
also have [P ] = [Q] iff there exists a finitely generated free module Rn such that
P ⊕ Rn ∼ = Q ⊕ Rn . In particular, [P ] = [0] = 0 iff P is stably free: there exists a
52 3. MODULES

finitely generated free module F such that P ⊕ F is free.


d) Show that M0 (R) is cancellative iff every stably free finitely generated projective
module is free.
e)* Find a ring R admitting a finitely generated projectve module which is stably
free but not free.
f ) Show that the mapping Rn 7→ [Rn ] induces an injective homomorphism of rings
Z → K0 (R). Define K̃0 (R) to be the quotient K0 (R)/Z. Show that if R is a PID
then K̃0 (R) = 0.
Exercise 3.34. Insert exercise here.

6. Injective modules
6.1. Basic equivalences.

Although we will have no use for them in the sequel of these notes, in both commu-
tative and (especially) homological algebra there is an important class of modules
“dual” to the projective modules. They are characterized as follows.
Proposition 3.19. For a module E over a ring R, the following are equivalent:
(ii) If ι : M → N is an injective R-module homomorphism and ϕ : M → E is
any homomorphism, there exists at least one extension of ϕ to a homomorphism
Φ : N → E.
(iii) If M ,→ N , the natural map Hom(N, E) → Hom(M, E) is surjective.
(iv) The (contravariant) functor Hom( , E) is exact.
(v) Any short exact sequence of R-modules
ι
0→E→M →N →0
splits: there exists an R-module map π : M → E such that π ◦ ι = 1E and thus an
internal direct sum decomposition M = ι(E) ⊕ ker(π) ∼ = E ⊕ N.
A module satisfying these equivalent conditions is called injective.
Exercise 3.35. Prove Proposition 3.19.
Exercise 3.36. Show: an R-module E is injective iff whenever E is a submod-
ule of a module M , E is a direct summand of M .
Notice that the set of equivalent conditions starts with (ii)! This is to facilitate
direct comparison to Proposition 3.10 on projective modules. Indeed, one should
check that each of the properties (ii) through (v) are duals of the corresponding
properties for projective modules: i.e., they are obtained by reversing all arrows.
The difficulty here with property (i) is that if one literally reverses the arrows in the
definition of free R-module to arrive at a “cofree” R-module, one gets a definition
which is unhelpfully strong: the “cofree R-module on a set X” does not exist when
#X > 1! This can be remedied by giving a more refined definition of cofree module.
For the sake of curiosity, we will give it later on in the exercises, but to the best
of my knowledge, cofree R-modules by any definition do not play the fundamental
role that free R-modules do.
Exercise 3.37. Show: every module over a field is injective.
Exercise 3.38. Show: Z is not an injective Z-module.
(Injectivity is the most important property of modules that is not necessarily satisfied
by free modules.)
6. INJECTIVE MODULES 53

Q
Exercise 3.39. Let {Mi }i∈I be any family of R-modules and put M = i∈I Mi .
Show that M is injective iff Mi is injective for all i ∈ I.
Exercise 3.40. For a ring R, show the following are equivalent:
(i) R is absolutely projective: every R-module is projective.
(ii) R is absolutely injective: every R-module is injective.
6.2. Baer’s Criterion.
Theorem 3.20. (Baer’s Criterion [Ba40]) For a module E over a ring R, the
following are equivalent:
(i) E is injective.
(ii) For every ideal nonzero I of R, every R-module map ϕ : I → E extends to an
R-module map Φ : R → E.
Proof. (i) =⇒ (ii): this is a special case of condition (ii) of Proposition 3.19:
take M = I, N = R.
(ii) =⇒ (i): Let M be an R-submodule of N and ϕ : M → E an R-module map.
We need to show that ϕ may be extended to N . Now the set P of pairs (N 0 , ϕ0 )
with M ⊂ N 0 ⊂ N and ϕ : N 0 → E a map extending ϕ is nonempty and has an
evident partial ordering, with respect to which the union of any chain of elements
in P is again an element of P. So by Zorn’s Lemma, there is a maximal element
ϕ0 : N 0 → E. Our task is to show that N 0 = N .
Assume not, and choose x ∈ N \ N 0 . Put
I = (N 0 : x) = {r ∈ R | rx ⊂ N 0 };
one checks immediately that I is an ideal of R (a generalization to modules of the
colon ideal we have encountered before). Consider the composite map
·x ϕ
I → N 0 → E;
by our hypothesis, this extends to a map ψ : R → E. Now put N 00 = hN 0 , xi and
define5 ϕ00 : N 00 → E by
ϕ00 (x0 + rx) = ϕ0 (x0 ) + ψ(r).
Thus ϕ00 is an extension of ϕ0 to a strictly larger submodule of N than N 0 , contra-
dicting maximality. 
Exercise 3.41. Verify that the map ϕ00 is well-defined.
6.3. Divisible modules.

Recall that a module M over a domain R is divisible if for all r ∈ R• the endomor-
phism r• : M → M, x 7→ rx, is surjective. Further, we define M to be uniquely
divisible if for all r ∈ R• , the endomorphism r• : M → M is a bijection.
Example 3.21. The Z-modules Q and Q/Z are divisible. Q is moreover uniquely
divisible but Q/Z is not.
Exercise 3.42. Show: a divisible module is uniquely divisible iff it is torsion-
free.
5Since N 00 need not be the direct sum of N 0 and hxi, one does need to check that ϕ00 is
well-defined; we ask the reader to do so in an exercise following the proof.
54 3. MODULES

Exercise 3.43. a) Show: a quotient of a divisible module is divisible.


b) Show: arbitrary direct sums and direct products of divisible modules are divisible.
Exercise 3.44. Let R be a domain with fraction field K.
a) Show: K is a uniquely divisible R-module.
b) Let M be any R-module. Show that the natural map M → M ⊗R K is injective
iff M is torsionfree.
c) Show: for any R-module M , M ⊗R K is uniquely divisible.
d) Show: K/R is divisible but not uniquely divisible.
Exercise 3.45. a) Show: a Z-module is uniquely divisible iff it can be endowed
with the compatible structure of a Q-module, and if so this Q-module structure is
unique.
b) Show that a Z-module M is a subgroup of a uniquely divisible divisible Z-module
iff it is torsionfree.
Exercise 3.46. For a domain R, show that the following are equivalent:
(i) There is a nonzero finitely generated divisible R-module.
(ii) R is a field.
Proposition 3.22. Let R be a domain and E an R-module.
a) If E is injective, it is divisible.
b) If E is torsionfree and divisible, it is injective.
c) If R is a PID and E is divisible, it is injective.
Proof. a) Let r ∈ R• . For x ∈ E, consider the R-module homomorphism
ϕ : rR → E given by r 7→ x. Since E is injective, this extends to an R-module map
ϕ : R → E. Then rϕ(1) = ϕ(r · 1) = ϕ(r) = x, so r• is surjective on E.
b) Let I be a nonzero ideal of R and ϕ : I → E be an R-module map. For each
a ∈ I • , there is a unique ea ∈ E such that ϕ(a) = aea . For b ∈ I • , we have
baea = bϕ(a) = ϕ(ba) = aϕ(b) = abeb ;
since E is torsionfree we conclude ea = eb = e, say. Thus we may extend ϕ to a
map Φ : R → E by Φ(r) = re. Thus E is injective by Baer’s Criterion.
c) As above it is enough to show that given a nonzero ideal I of R, every homo-
morphism ϕ : I → E extends to a homomorphism R → E. Since R is a PID, we
may write I = xR for x ∈ R• . Then, as in part a), one checks that ϕ extends to Φ
iff multiplication by x is surjective on M , which it is since M is divisible. 
By combining Proposition 3.22 with Exercise 3.42, we are able to show an important
special case of the desired fact that every R-module can be realized as a submodule
of an injective module. Namely, if M is a torsionfree module over a domain R, then
M is a submodule of the uniquely divisible – hence injective – module M ⊗R K.
Exercise 3.47. Let n ∈ Z+ .
a) Show: as a Z-module Z/nZ is not divisible hence not injective.
b) Show: as a Z/nZ module Z/nZ is divisible iff n is a prime number.
c) Show: Z/nZ is always injective as a Z/nZ-module.
Exercise 3.48. Let R = Z[t] and let K be its fraction field. Show: the R-
module K/R is divisible but not injective.
Exercise 3.49. Let R be a domain with fraction field K.
a) If R = K, show: all R-modules are both injective and projective.
b) If R 6= K, show: the only R-module that is both projective and injective is 0.
6. INJECTIVE MODULES 55

6.4. Enough injectives.

The idea of this section is to pursue the dual version of the statement “Every
R-module is a quotient of a projective module”: namely we wish to show that
every R-module is a submodule of an injective module. This is a good example
of a statement which remains true upon dualization but becomes more elaborate
to show. The projective version is almost obvious: indeed, we have the stronger
result that every module is a quotient of a free module, and – as we have seen – to
realize M as a quotent of a free R-module is equivalent to simply choosing a set of
generators for M . (But again, if we choose the most obvious definition of “cofree”,
then this statement will be false.)

Let k be a ring, R a k-algebra, M an R-module and N a k-module. Consider


the commutative group Homk (M, N ). We may endow it with the structure of an
R-module as follows: for r ∈ R and f ∈ HomZ (M, N ), (rf )(x) := f (rx).

Consider the special case k = Z and N = Q/Z of the above construction. It


gives HomZ (M, Q/Z) the structure of an R-module, which we denote by M ∗ and
call the Pontrjagin dual of M .6 Because Q/Z is an injective Z-module, the (con-
travariant) functor M 7→ M ∗ – or in other words HomZ ( , Q/Z) – is exact.7 In
particular, if f : M → N is an R-module map, then f injective implies f ∗ surjec-
tive and f surjective implies f ∗ injective.

As is often the case for “duals”, we have a natural map M → M ∗∗ : namely


x 7→ (f 7→ f (x)).
Lemma 3.23. For any R-module M , the natural map ΨM : M → M ∗∗ is
injective.
Proof. Seeking a contradiction, let x ∈ M • be such that Ψ(x) = 0. Unpacking
the definition, this means that for all f ∈ HomZ (M, Q/Z), f (x) = 0. But since Q/Z
is an injective Z-module, it suffices to find a nontrivial homomorphism Zx → Q/Z,
and this is easy: if x has finite order n > 1, we may map x to n1 , whereas if x has
infinite order we may map it to any nonzero element of Q/Z. 
Lemma 3.24. Every Z-module M can be embedded into an injective Z-module.
L
Proof. Let I ⊂ M be a generating set andL let i∈I Z → M be the cor-
responding surjection, with kernel K, so M ∼
= ( Z)/K. The natural map
L L L i∈I
i∈I Z ,→ i∈I Q induces an injection M ,→ ( i∈I Q)/K, and the latter Z-
module is divisible, hence injective since Z is a PID. 
Lemma 3.25. (Injective Production Lemma) Let R be a k-algebra, E an injec-
tive k-module and F a free R-module. Then Homk (F, E) is an injective R-module.
Proof. We will show that the functor HomR ( , Homk (F, E)) is exact. For any
R-module M , the adjointness of ⊗ and Hom gives
HomR (M, Homk (F, E)) = Homk (F ⊗R M, E)
6Recall that the notation M ∨ has already been taken: this is the linear dual Hom (M, R).
R
7Here we are using the (obvious) fact that a sequence of R-modules is exact iff it is exact
when viewed merely as a sequence of Z-modules.
56 3. MODULES

so we may look at the functor M 7→ Homk (F ⊗R M, E) instead. This is the


composition of the functor M 7→ F ⊗R M with the functor N 7→ Homk (N, E). But
both functors are exact – in the former case a moment’s thought shows this to be
true, and the latter case is one of our defining properties of injective modules. 
Remark: Soon enough we will define a flat R-module to be an R-module N such
that the functor M 7→ M ⊗R N is exact. Then Lemma 3.25 can be rephrased with
the hypothesis that F is a flat R-module, and (since as we have just seen, free
R-modules are flat) this gives a somewhat more general result.
Theorem 3.26. Every R-module can be embedded into an injective R-module.
Proof. Let M be an R-module. Viewing M as a Z-module, by Lemma 3.24
there is an injective Z-module E1 and a Z-module map ϕ1 : M ,→ E1 . Further, by
Lemma 3.25, HomZ (R, E1 ) is an injective R-module. Now consider the R-module
map
ϕ : M → HomZ (R, E1 ), x 7→ (r 7→ ϕ1 (rx)).
We claim that ϕ is a monomorphism into the injective R-module HomZ (R, E1 ).
Indeed, if ϕ(x) = 0 then for all r ∈ R, ϕ1 (rx) = 0. In particular ϕ1 (x) = 0, so since
ϕ1 is a monomorphism, we conclude x = 0. 
Exercise 3.50. We say a Z-module is cofree if it is of the form F ∗ for a
free Z-module F . Then the proof of Lemma 3.24 gives the stronger statement that
every Z-module can be embedded into a cofree Z-module. Formulate a definition of
cofree R-module so that the proof of Theorem 3.26 gives the stronger statement
that every R-module can be embedded into a cofree R-module. (Hint: remember to
pay attention to the difference between direct sums and direct products.)
6.5. Essential extensions and injective envelopes.

The results of this section are all due to B. Eckmann and A. Schopf [ES53].
Proposition 3.27. For R modules M ⊂ N , the following are equivalent:
(i) If X is any nonzero R-submodule of N , then X ∩ M is nonzero.
(ii) If x ∈ N • , there exists r ∈ R such that rx ∈ M • .
(iii) If ϕ : N → Y is an R-module map, then ϕ is injective iff ϕ|M is injective.
An extension M ⊂ N satisfying these equivalent conditions is called essential.
Proof. (i) =⇒ (ii): Apply (i) with X = hxi.
(ii) =⇒ (iii): Assuming (ii), let ϕ : N → Y be a homomorphism with ϕ|M is
injective. It is enough to show that ϕ is injective. Seeking a contradiction, let
x ∈ N • be such that ϕ(x) = 0. By (ii), there exists r ∈ R such that rx ∈ M • . But
then by assumption rϕ(x) = ϕ(rx) 6= 0, so ϕ(x) 6= 0, contradiction.
(iii) =⇒ (i): We go by contraposition. Suppose there exists a nonzero submodule
X of N such that X ∩ M = 0. Then the map ϕ : N → N/X is not an injection but
its restriction to M is an injection. 
Proposition 3.28. (Tower Property of Essential Extensions) Let L ⊂ M ⊂ N
be R-modules. Then L ⊂ N is an essential extension iff L ⊂ M and M ⊂ N are
both essential extensions.
Proof. Suppose first that L ⊂ N is an essential extension. Then for any
nonzero submodule X of N , X ∩ L 6= 0. In particular this holds for X ⊂ M , so
6. INJECTIVE MODULES 57

L ⊂ M is essential. Moreover, since L ⊂ M , X ∩ L 6= 0 implies X ∩ M 6= 0, so


M ⊂ N is essential. Conversely, suppose L ⊂ M and M ⊂ N are both essential,
and let X be a nonzero submodule of N . Then X ∩ M is a nonzero submodule of
M and thus (X ∩ M ) ∩ L = X ∩ L is a nonzero submodule of L. So L ⊂ N is
essential. 
So why are we talking about essential extensions when we are supposed to be talking
about injective modules? The following result explains the connection.
Theorem 3.29. For an R-module M , the following are equivalent:
(i) M is injective.
(ii) M has no proper essential extensions: i.e., if M ⊂ N is an essential extension,
then M = N .
Proof. (i) =⇒ (ii): Let M be injective and M ( N . Then M is a direct
summand of N : there exists M 0 such that M ⊕ M 0 = N . Thus M has zero
intersection with M 0 , and by criterion (ii) of Proposition 3.27, we must have M 0 = 0
and thus M = N .
(ii) =⇒ (i): It suffices to show: if N is an R-module and M ⊂ N , then M is a
direct summand of N . Now consider the family of submodules M 0 of N with the
property that M ∩ M 0 = 0. This family is partially ordered by inclusion, nonempty,
and closed under unions of chains, so by Zorn’s Lemma there exists a maximal such
element M 0 . Now consider the extension M ,→ N/M 0 : we claim it is essential.
Indeed, if not, there exists x ∈ N \ M 0 such that hM 0 , xi ∩ M = 0, contradicting
maximality of M 0 . But by hypothesis, M has no proper essential extensions: thus
M = N/M 0 , i.e., M ⊕ M 0 = N and M is a direct summand of N . 
We say that an extension M ⊂ N is maximal essential if it is essential and
there is no proper extension N 0 of N such that M ⊂ N 0 is essential. Combining
Proposition 3.28 and Theorem 3.29 yields the following important result.
Theorem 3.30. For an essential extension M ⊂ N of R-modules, the following
are equivalent:
(i) M ⊂ N is maximal essential.
(ii) N is injective.
Exercise 3.51. To be sure you’re following along, prove Theorem 3.30.
Once again we have a purpose in life – or at least, this subsection of it – we would like
to show that every R-module admits a maximal essential extension and that such
extensions are unique up to isomorphism over M . Moreover, a plausible strategy
of proof is the following: let M be an R-module. By Theorem 3.26 there exists an
extension M ⊂ E with E injective. Certainly this extension need not be essential,
but we may seek to construct within it a maximal essential subextension N and
then hope to show that M ⊂ E 0 is injective.
Theorem 3.31. Let M be an R-module and M ⊂ E an extension with E
injective. Let P be the set of all essential subextensions N of M ⊂ E. Then:
a) P contains at least one maximal element.
b) Every maximal element E 0 of P is injective.
Proof. The proof of part a) is the usual Zorn’s Lemma argument: what we
need to check is that the union N of any chain {Ni } of essential subextensions
58 3. MODULES

is again an essential subextension. Suppose for a contradiction that there exists


a nonzero submodule X of N such that X ∩ M = 0. Choose x ∈ X • and put
X 0 = hxi. Then X 0 ⊂ Ni for some i and X 0 ∩ M ⊂ X ∩ M = 0, contradicting the
essentialness (?!) of the extension M ⊂ Ni .
Now let E 0 be a maximal essential subextension of M ⊂ E. We need to show
that M ⊂ E 0 is actually a maximal essential extension: so suppose there is an
essential extension E 0 ⊂ N . Let ι : M ⊂ E 0 ⊂ N be the composite map. It is
a monomorphism, so by the injectivity of E the injection M ⊂ E extends to a
homomorphism ϕ : N → E. But ϕ|M is an injection and M ⊂ N is an essential
extension, so by condition (iii) of Proposition 3.27 this implies that ϕ itself is an
injection. By maximality of E 0 among essential subextensions of M ⊂ E we must
have E 0 = N . 
For an R-module M , we say that an extension M ⊂ E is an injective envelope
(other common name: injective hull) of M if M ⊂ E is a maximal essential
extension; equivalently, an essential extension with E injective. Thus Theorem
3.31 shows that any R-module admits an injective envelope.
Proposition 3.32. Let R be a domain with fraction field K. Then R ⊂ K is
an injective envelope of R.
Exercise 3.52. Prove Proposition 3.32.
(Suggestion: use the relationship between injective modules and divisible modules.)
Exercise 3.53. More generally, let M be a torsionfree module over a domain
R. Show that M ⊂ M ⊗R K is an injective envelope of M .
Let us touch up our characterization of injective envelopes a bit.
Proposition 3.33. (Equivalent Properties of an Injective Envelope) For an
extension M ⊂ E of R-modules, the following are equivalent:
(i) M ⊂ E is a maximal essential extension.
(ii) M ⊂ E is essential and E is injective.
(iii) E is minimal injective over M : there does not exist any proper subextension
M ⊂ E 0 ⊂ E with E 0 injective.
Proof. We have already seen that (i) ⇐⇒ (ii).
(ii) =⇒ (iii): Assume that E is injective and E 0 is an injective subextension of
M ⊂ E. Since E 0 is injective, there exists N ⊂ E such that E 0 ⊕ N = E. Moreover,
M ∩ N ⊂ E 0 ∩ N = 0, so M ∩ N = 0. Since M ⊂ E is essential, we must have
N = 0, i.e., E 0 = E.
(iii) =⇒ (ii): Suppose that M ⊂ E is minimal injective. The proof of Theorem
3.31 gives us a subextension E 0 of M ⊂ E such that E 0 is injective and M ⊂ E 0 is
essential. Thus by minimality E = E 0 , i.e., M ⊂ E is essential. 
Theorem 3.34. (Uniqueness of Injective Envelopes) Let M be an R-module
and let ι1 : M ⊂ E1 , ι2 : M ⊂ E2 be two injective envelopes of M . Then E1 and
E2 are isomorphic as R-module extensions of M : i.e., there exists an R-module
isomorphism Φ : E1 → E2 such that Φ ◦ ι1 = ι2 .
Proof. Since ι1 : M → E1 is a monomorphism and E2 is injective, the map ι2 :
M → E2 extends to a map Φ : E1 → E2 such that Φ ◦ ι1 = ι2 . Since the restriction
of Φ to the essential submodule M is a monomorphism, so is Φ. The image Φ(E1 )
7. FLAT MODULES 59

is an essential subextension of M ⊂ E2 , so by condition (iii) of Proposition 3.33 we


must have E2 = Φ(E1 ). Thus Φ : E1 → E2 is an isomorphism. 
In view of Theorem 3.34, it is reasonable to speak of “the” injective envelope of
M and denote it by M → E(M ). Reasonable, that is, but not ideal: it is not
true that any two injective envelopes are canonically isomorphic.8 Otherwise put,
formation of the injective envelope is not functorial. For more on this in a more
general category-theoretic context, see [AHRT].
Exercise 3.54. Let M be a submodule of an injective module E. Show: E
contains an isomorphic copy of the injective envelope E(M ).
Exercise 3.55. If M ⊂ N is an essential extension of modules, then E(M ) =
E(N ).

7. Flat modules
Suppose we have a short exact sequence
0 → M 0 → M → M 00 → 0
of R-modules. If N is any R-module, we can tensor each element of the sequence
with N , getting by functoriality maps
0 → M 0 ⊗ N → M ⊗ N → M 00 ⊗ → 0.
Unfortunately this new sequence need not be exact. It is easy to see that it is right
exact: that is, the piece of the sequence
M 0 ⊗ N → M ⊗ N → M 00 ⊗ N → 0
remains exact. This follows because of the canonical “adjunction” isomorphism
Hom(M ⊗ N, P ) = Hom(M, Hom(N, P ))
and the left-exactness of the sequence Hom( , Y ) for all R-modules Y . However,
tensoring an injection need not give an injection. Indeed, consider the exact se-
quence
[2]
0 → Z → Z.
If we tensor this with Z/2Z, we get a sequence
[2]
0 → Z/2Z → Z/2Z,
but now the map Z ⊗ Z/2Z → Z ⊗ Z/2Z takes n ⊗ i → (2n ⊗ i) = n ⊗ 2i = 0, so is
not injective.

Definition: A module M over a ring R is flat if the functor N 7→ N ⊗R M is


exact. This means, equivalently, that if M ,→ M 0 then M ⊗ N ,→ M 0 ⊗ N , or also
that tensoring a short exact sequence with M gives a short exact sequence.

It will probably seem unlikely at first, but in fact this is one of the most important
and useful properties of an R-module.

So, which R-modules are flat?


8The situation here is the same as for “the” splitting field of an algebraic field extension or
“the” algebraic closure of a field.
60 3. MODULES

Proposition 3.35. Let {Mi }i∈I be a family of R-modules. The following are
equivalent:
(i) For all i, Mi is flat. L
(ii) The direct sum M = i Mi is flat.
Exercise 3.56. Prove Proposition 3.35.
Proposition 3.36. Let R be a domain. Then flat R-modules are torsionfree.
Proof. We will prove the contrapositive. Suppose that 0 6= m ∈ R[tors], and
let 0 6= r ∈ R be such that rm = 0 Since R is a domain, we have a short exact
sequence
[r]
0 → R → R → R/rR → 0
and tensoring it with M gives
[r]
0 → M → M → M/rM → 0,
but since rm = 0 the first map is not injective. 
Example 3.37. Let k be a field, and let R := k[x, y], the polynomial ring in
two indeterminates over k. Let I := hx, yi; then I is a maximal ideal such that
R/I = k. Like every ideal of a domain, I is a torsionfree module. We claim that I
is not flat.9 It suffices to show that the map
ϕ : I ⊗ I → I ⊗R R = I
obtained by tensoring the injection I ,→ R with I is not an injection. This map is
nothing else Pn map from the R-bilinear multiplication map I ×I → I:
Pn than the induced
it sends i=1 ai ⊗ bi to i=1 ai yi . Evidently the element
θ := x ⊗ y − y ⊗ x
lies in the kernel of ϕ; the crux of the matter is to show that θ 6= 0. For this, let
Dx : k[x, y] → k = k[x, y]/I
by taking the partial derivative with respect to x and then evaluating at (0, 0) and
let
Dy : k[x, y] → k = k[x, y]/I
by taking the partial derivative with respect to y and then evaluating at (0, 0). Fi-
nally, put
D : I × I → k, (a, b) 7→ Dx (a)Dy (b).
Then D is R-bilinear: e.g. for a, b ∈ I = hx, yi and c, d ∈ k[x, y] we have
  
∂a ∂c ∂b ∂d
D(ca, db) = Dx (ca)Dy (db) = c +a d +b + hx, yi
∂x ∂x ∂y ∂y
∂a ∂b
= cd + hx, yi.
∂x ∂y
So D factors through an R-linear map D : I ⊗ I → k. Since
D(θ) = D(x ⊗ y − y ⊗ x) = D((x, y)) − D((y, x)) = 1 − 0 = 1,
it follows that θ 6= 0.
9Much later we will learn that because I is a prime ideal of height greater than 1 in a
Noetherian ring, it cannot be flat.
8. NAKAYAMA’S LEMMA 61

Proposition 3.38. Projective R-modules are flat.


Proof. A projective R-module is a module P such that there exists P 0 with
P ⊕P 0 ∼= F a free module. Therefore, by Proposition 3.35, it is enough to show that
free modules are flat. By abuse of notation, we will abbreviate the infinite direct
sum of d copies of R as Rd . Since for any R-module M we have M ⊗R Rd = M d ,
it follows that tensoring a short exact sequence
0 → M 0 → M → M 00 → 0
with F = Rd just yields
0 → (M 0 )d → (M )d → (M 00 )d → 0.
This is still exact. 

8. Nakayama’s Lemma
8.1. Nakayama’s Lemma.
Proposition 3.39. Let M be a finitely generated R-module, I an ideal of R,
and ϕ be an R-endomorphism of M such that ϕ(M ) ⊂ IM . Then ϕ satisfies an
equation of the form
ϕn + an−1 ϕn−1 + . . . + a1 ϕ + a0 = 0,
with ai ∈ I.
Proof. Let x1 , . . . , xn be a set of generators
P for M as an R-module. Since
each ϕ(xi ) ∈ IM , we may write ϕ(xi ) = j aij xj , with aij ∈ I. Equivalently, for
all i,
Xn
(δij ϕ − aij )xj = 0.
j=1

By multiplying on the left by the adjoint of the matrix M = (δij ϕ − aij ), we get
that det(δij ϕ − aij ) kills each xi , hence is the zero endomorphism of M . Expanding
out the determinant gives the desired polynomial relation satisfied by ϕ. 

Exercise 3.57. Some refer to Proposition 3.39 as the Cayley-Hamilton Theo-


rem. Discuss.
Theorem 3.40 (Nakayama’s Lemma). Let R be a ring, J an ideal of R, and
M a finitely generated R-module such that JM = M .
a) There is x ∈ R such that x ≡ 1 (mod J) and xM = {0}.
b) If moreover J is contained in every maximal ideal of R, then M = {0}.
Proof. Applying Proposition 3.39 to the identity endomorphism ϕ = 1M gives
a1 , . . . , an−1 ∈ J such that for x := 1 + a1 + . . . + an−1 , we have xM = {0} and
x ≡ 1 (mod J), proving part a). If moreover J lies in every maximal ideal m of R,
then x ≡ 1 (mod m) for all maximal ideals m, hence x lies in no maximal ideal of
R. Therefore x is a unit and multiplying xM = {0} by x−1 gives M = {0}. 

Corollary 3.41. Let R be a ring, J an ideal of R which is contained in every


maximal ideal of R, M a finitely generated R-module and N a submodule of M
such that JM + N = M . Then M = N .
62 3. MODULES

Proof. We have J(M/N ) = (JM + N )/N = M/N . Applying Nakayama’s


Lemma to the finitely generated module M/N , we conclude M/N = 0, i.e., N =
M. 

Corollary 3.42. Let R be a ring, J an ideal of R which is contained in every


maximal ideal of R, and M a finitely generated R-module. Let x1 , . . . , xn ∈ M be
such that their images in M/JM span M/JM as an R/J-module. Then the xi ’s
span M .
Proof. Let N = hx1 , . . . , xn iR , and apply Corollary 3.41. 

Corollary 3.43. Let R be a ring and J an ideal which is contained in every


maximal ideal of R. Let M and N be R-modules, with N finitely generated, and let
u : M → N be an R-module map. Suppose that the map uJ : M/JM → N/JN is
surjective. Then u is surjective.
Proof. Apply Nakayama’s Lemma to J and N/M . 

Recall that an element x in a ring R such that x2 = x is called idempotent.


Similarly, an ideal I of R such that I 2 = I is called idempotent.
Exercise 3.58. Let R be a ring and I an ideal of R.
a) Suppose I = (e) for an idempotent element e. Show that I is idempotent.
b) Give an example of a nonidempotent x such that (x) is idempotent.
c) Is every idempotent ideal generated by some idempotent element?
Corollary 3.44. Let R be any ring and I a finitely generated idempotent
ideal of R. Then there is an idempotent e ∈ R such that I = (e). In particular,
in a Noetherian ring every idempotent ideal is generated by a single idempotent
element.
Exercise 3.59. Prove Corollary 3.44. (Hint: apply Theorem 3.40!)
8.2. Hopfian modules.

A group G is Hopfian if every surjective group homomorphism f : G → G is


an isomorphism – equivalently, G is not isomorphic to any of its proper quotients.
This concept has some currency in combinatorial and geometric group theory.
Some basic examples: any finite group is certainly Hopfian. A free group is Hopfian
iff it is finitely generated, and more generally a finitely generated residually
Q∞ finite
group is Hopfian. An obvious example of a non-Hopfian group is i=1 G for any
nontrivial group G. A more interesting example is the Baumslag-Solitar group
B(2, 3) = hx, y | yx2 y −1 = x3 i.
More generally, let C be a concrete category: that is, Ob C is a class of sets and
for all X, Y ∈ Ob C, HomC (X, Y ) ⊂ HomSet (X, Y ), i.e., the morphisms between X
and Y are certain functions from X to Y . We may define an object X in C to be
Hopfian if every surjective endomorphism of X is an isomorphism.
Exercise 3.60. a) (C. LaRue) Show: any finite object in a concrete cat-
egory is Hopfian.
b) In the category of sets, the Hopfian objects are precisely the finite sets.
8. NAKAYAMA’S LEMMA 63

Remark: Our discussion of “Hopfian objects” in categories more general than R-


Mod is not particularly serious or well thought out. So far as I know there is
not a completely agreed upon definition of a Hopfian object, but Martin Branden-
burg has suggested (instead) the following: X ∈ C is Hopfian if every extremal
epimorphism X → X is an isomorphism.
Theorem 3.45. Let R be a ring and M a finitely generated R-module. Then
M is a Hopfian object in the category of R-modules.
Proof. ([M, p. 9]) Let f : M → M be a surjective R-map. We show f is
injective.
There is a unique R[t]-module structure on M extending the given R-module
structure and such that for all m ∈M, tm = f (m). Let I = tR[t]. By hypothesis
IM = M , so by Nakayama’s Lemma there exists P (t) ∈ R[t] such that (1 +
P (t)t)M = 0. Let y ∈ ker f . Then
0 = (1 + P (t)t)y = y + P (t)f (y) = y + P (t)0 = y.
So f is injective. 
Exercise 3.61. Show: (Q, +) is a Hopfian Z-module which is not finitely gen-
erated.
Exercise 3.62. Do there exist Hopfian Z-modules of all cardinalities? (An
affirmative answer was claim in [Ba62], but it was announced in [Ba63] that the
construction is not valid. So far as I know the problem remains open lo these many
years later.)
8.3. A variant.

The results of this section are taken from [DM71, §I.1].


Theorem 3.46 (Nakayama’s Lemma). Let R be a ring, J an ideal of R and
M a finitely generated R-module. the following are equivalent:
(i) J + ann M = R.
(ii) JM = M .
Proof. (i) =⇒ (ii): The annihilator of M/JM contains J and ann M , so it
contains J + ann M = R. This means that M/JM = 0 and JM = M .
(ii) =⇒ (i): Conversely, suppose M = hm1 , . . . , mn i. For 1 ≤ i ≤ n, put
Mi = hmi , . . . , mn i and Mn+1 = 0. We claim that for all 1 ≤ i ≤ n + 1 there exists
ai ∈ J with (1 − ai )M ⊂ Mi , and we will prove this by induction on n. We may
take a1 = 0. Having chosen a1 , . . . , ai , we have
(1 − ai )M = (1 − ai )JM = J(1 − ai )M ⊂ Mi ,
so there exist aij ∈ J such that
n
X
(1 − ai )mj = aij mj ,
j=i
or
(1 − ai − aii )mi ∈ Mi+1 .
Thus
1 − (2ai + aii − a2i − ai ai i) M = (1−ai )(1−ai −aii )M ⊂ (1−ai −aii )Mi ⊂ Mi+1 ,

64 3. MODULES

and we may take


ai+1 = 2ai + aii − a2i − ai ai i.
So there is an ∈ J such that 1 − an ∈ ann M , and thus 1 ∈ J + ann M . 

Exercise 3.63. Explain why Theorem 3.46 is an equivalent (but nicer?) refor-
mulation of Theorem 3.40a).
Corollary 3.47. Let M be a finitely generated R-module such that mM = M
for all maximal ideals of R. Then M = 0.
Exercise 3.64. Prove Corollary 3.47.
For an R-module M , we define its trace ideal to be the ideal T (M ) of R generated
by all the images f (M ) of R-module maps f ∈ R∨ = HomR (M, R).
Theorem 3.48. Let P be a finitely generated projective R-module. Then R
splits as a direct product of rings:
R = T (P ) ⊕ ann P.
Proof. Step 1: We show that T (P ) and ann P are comaximal ideals of R, i.e.,
T (P ) + ann P = R. By the Dual Basis Lemma (Proposition 3.12) and the following
exercise, there exist x1 , . . . , xn ∈ PPand f1 , . . . , fn ∈ P ∨ = HomR (P, R) forming
n
a dual basis: for all x ∈ P , x = i=1 fi (x)xi . By its very definition we have
fi (x) ∈ T (P ) for all i and x, hence T (P )P = P . By the Generalized Nakayama’s
Lemma (Lemma 3.46) we have T (P ) + ann P = R.
For any a ∈ ann P , f ∈ P ∨ and x ∈ P we have af (x) = f (ax) = f (0) = 0: thus
T (P ) ∩ ann P = 0. By comaximality, T (P ) ∩ ann P = 0 and the sum is direct. 

Corollary 3.49. A nonzero finitely generated projective module over a con-


nected ring R (i.e., without idempotents other than 0 and 1) is faithful.
Exercise 3.65. Prove Corollary 3.49.
8.4. Applications to modules over local rings.
Lemma 3.50. Let R be a ring and J an ideal which is contained in every
maximal ideal of R, and let M be a finitely presented R-module. Suppose that:
(i) M/JM is a free R/J-module, and
(ii) The canonical map J ⊗R M → JM is injective.
Then M is a free R-module.
Proof. We may choose a family {xi }i∈I of elements of M such that the images
in M/JM give a R/J-basis. (Since M is finitely generated over R, M/JM is finitely
generated over R/J, so the index
L set I is necessarily finite.) Consider the finitely
generated free R-module L = i∈I R, with canonical basis {ei }. Let u : L → M be
the unique R-linear mapping each ei to xi , and let K = ker(u). Since M is finitely
presented, by Proposition 3.6 K is finitely generated. We have a commutative
diagram with exact rows:
J ⊗K →J ⊗L→J ⊗M →0

0 → K → L → M → 0,
9. ORDINAL FILTRATIONS AND APPLICATIONS 65

where each vertical map – a : J ⊗ K → K, b : J ⊗ L → L, c : J ⊗ M → M – is the


natural multiplication map. Our hypothesis is that the the map J ⊗R M → JM is
injective, so by the Snake Lemma we get an exact sequence
u
0 → coker(a) → coker(b) → coker(c).
Now observe that coker(b) = (R/J) ⊗R L and coker(c) = (R/J) ⊗R M , and by
definition the mapping u : L → M gives, upon passage to the quotient modulo J, a
mapping from one R/J-module basis to another. So u is an isomorphism and thus
coker(a) = 0, i.e., K/JK = 0. By Nakayama’s Lemma we conclude K = 0, i.e., u
gives an isomorphism from the free module L to M , so M is free. 
We can now prove the following result, which is one that we will build upon in our
future studies of modules over commutative rings.
Theorem 3.51. Let R be a ring with a unique maximal ideal m – i.e., a local
ring. For a finitely presented R-module M , the following are equivalent:
(i) M is free.
(ii) M is projective.
(iii) M is flat.
(iv) The natural map m ⊗R M → mM is an injection.
Proof. Each of the implications (i) =⇒ (ii) =⇒ (iii) =⇒ (iv) is immediate.
Assume (iv). Then, since m is maximal, R/m is a field, so every R/m-module is
free. Therefore Lemma 3.50 applies to complete the proof. 

9. Ordinal Filtrations and Applications


9.1. The Transfinite Dévissage Lemma.

Let M be an R-module. By an ordinal filtration on M we mean an ordinal


number α and for each i ≤ α a submodule Mi of M satisfying all of the following:
(OF1) M0 = 0, Mα = M .
(OF2) For all i, j ∈ α + 1, i ≤ j =⇒ Mi S
⊂ Mj .
(OF3) For all limit ordinals i ≤ α, Mi = j<i Mj .

So for instance, taking α = ω = {1, 2, 3, . . .} the first infinite ordinal, we recover


the usual notion of an exhaustive
S filtration by submodules Mn , with the additional
convention that Mω = n∈ω Mn .

For i < α, we call Mi+1 /Mi the ith successive quotient. If for a class C of
R-modules each successive quotient lies in C, we say the filtration is of class C.
L
Define the associated graded module Gr(M ) = i<α Mi+1 /Mi .
Lemma 3.52. (Transfinite Dévissage Lemma) Let M be an R-module and
{Mi }i≤α an ordinal filtration of M .
a) Suppose we make the following hypothesis:
(DS) For all i < α the submodule Mi is a direct summand of Mi+1 . Then
M∼
M
= Gr(M ) = Mi+1 /Mi .
i<α
66 3. MODULES

b) Hypothesis (DS) holds if each successive quotient Mi+1 /Mi is projective.


c) Hypothesis (DS) holds if each Mi is injective.
Exercise 3.66. Prove Lemma 3.52. (Hint: use transfinite induction.)
Corollary 3.53. For an R-module M , the following are equivalent:
(i) M is free.
(ii) M admits an ordinal filtration with successive quotients isomorphic to R.
(iii) M admits an ordinal filtration with free successive quotients.
Proof. (i) =⇒ (ii): If M is free, then M = ∼ L R. By the Well-Ordering
i∈I
, I is in bijection with an ordinal α, so we may write M ∼
Principle10L
L
= i<α R, and
put Mi = j<i R.
(ii) =⇒ (iii) is immediate.
(iii) =⇒ (i) follows from Lemma 3.52 since free modules are projective. 
9.2. Hereditary and semihereditary rings.

An R-module is hereditary if every submodule is projective. (In particular a


hereditary module is projective, and thus the property of being projective is “in-
herited” by its submodules.) We say that a ring R is hereditary if R is a hereditary
R-module, or equivalently every ideal of R is projective as an R-module.
Exercise 3.67. a) Show: every submodule of a hereditary module is hereditary.
b) Show: the zero module is hereditary.
c) Show: there are nonzero rings R for which the only hereditary R-module is the
zero module.
Example 3.54. A PID is a hereditary ring. Indeed, any nonzero ideal of a
PID R is isomorphic as an R-module to R.
Theorem 3.55.
a) LFor i ∈ I, let Mi be a hereditary R-module, and let N be an R-submodule
of L i∈I Mi . Then there is for all i ∈ I an R-submodule Hi of Mi such that
N∼ = i∈I Hi .
b) For R hereditary and M an R-module, the following are equivalent:
(i) M is isomorphic to to a direct sum of ideals of R.
(ii) M can be embedded in a free R-module.
(iii) M can be embedded in a projective R-module.
L
Proof. a) Let M = i∈I be a direct sum of hereditary modules, and let N
be an R-submodule of M . By the Well-Ordering Principle there is a bijection
L from
I to some ordinal α, and L
without loss of generality we may assume M = i∈α Mi .
For j ∈ α+ , put Pj = i<j + Mi , so that {Mj } is an ordinal-indexed chain of
R-submodules of M with final element Pα = M . For each j ∈ α+ , put
N j = N ∩ Pj ,
so {Nj } is an ordinal filtration on N with Nα = N . Moreover, for all i ∈ α we have
Ni = Ni+1 ∩ Pi and thus
Ni+1 /Ni = Ni /(Ni+1 ∩ Pi ) ∼
= (Ni+1 + Pi )/Pi .
10This set-theoretic axiom is equivalent to the Axiom of Choice and also to Zorn’s Lemma.
Our running convention in these notes is to freely use these axioms when necessary.
9. ORDINAL FILTRATIONS AND APPLICATIONS 67

Thus Ni+1 /Ni is isomorphic to a submodule Hi of Pi+1 /Pi ∼ = Mi . For i ∈ I, since


Mi is hereditary, so is Hi ∼
= Ni+1 /Ni . In particular each Ni+1 /Ni is projective, and
the Transfinite Dévissage Lemma (Lemma 3.52) applies to show that
N∼ Ni+1 /Ni ∼
M M
= Gr N = = Hi .
i<α i∈I

b) (i) =⇒ (ii): Since an ideal is precisely an R-submodule of a free R-module of


rank 1, this holds over any ring.
(ii) ⇐⇒ (iii): Since every projective module is a direct summand of a free R-
module, this also holds over any ring.
(ii) =⇒ (i): This follows from part a). 
L
Corollary 3.56. Let {Mi }i∈I be a family of R-modules. Then M = i∈I Mi
is hereditary iff Mi is hereditary for all i.
Proof. Suppose each Mi is hereditary, and let N be a submodule of M .
By Theorem 3.55a), there is for all i ∈ I an R-submodule Hi of Mi such that
N ∼
L
= i∈I Hi . Each Hi , being a submodule of the hereditary module Mi , is itself
hereditary, hence projective. Thus N is a direct sum of projective modules, hence
projective. Conversely, if M is hereditary, so are all of its submodules Mi . 
Lemma 3.57. a) (Checking Projectivity With Injectives) Let P be an R-module
such that: for every injective module I, surjection q : I → Q and module map
f : P → Q, there is F : P → I such that q ◦ F = f . Then P is projective.
b) (Checking Injectivity With Projectives) Let I be an R-module such that: for every
projective R-module P , injection ι : S → P and module map f : S → I, there is
F : P → I such that F ◦ ι = f . Then I is injective.
ι τ
Proof. a) Let 0 → A0 → A → A00 → 0 be a short exact sequence of R-
modules, and let f : P → A00 be a module map. Let σ : A → I be an embedding
into an injective module, and consider the following commutative diagram with
exact rows:
ι τ
0 −−−−→ A0 −−−−→ A −−−−→ A00 −−−−→ 0

σ
y
σ◦ι q
0 −−−−→ A0 −−−−→ I −−−−→ Q −−−−→ 0
Step 1: We claim there is ρ : A00 → Q making the diagram commute.
Proof: This is a routine diagram chase: choose y ∈ A00 , lift to x in A, and put
ρ(y) = (q ◦ σ)(x). Let us check that this is well-defined: if we chose a different lift
x0 in A, then x − x0 ∈ A0 , so (q ◦ σ)(x − x0 ) = 0.
Step 2: By hypothesis, the map ρ ◦ f : P → Q can be lifted to G : P → I. To
complete the proof it suffices to show G(P ) ⊂ σ(A). To see this, let x ∈ P and
choose a ∈ A such that τ (a) = f (x). Then
q(G(x)) = ρ(f (x)) = ρ(τ (a)) = q(σ(a)),
so G(x) − σ(a) ∈ Ker q = Im(σ ◦ q). That is, there is a0 ∈ A0 such that σ(ι(a0 )) =
G(x) − σ(a), so
G(x) = σ(ι(a0 ) + a) ∈ σ(A).
b) This is the dual version of part a), i.e., obtained by reversing all the arrows. The
above proof also dualizes, as we leave it to the reader to check. 
68 3. MODULES

Corollary 3.58. (Cartan-Eilenberg)


For a ring R, the following are equivalent:
(i) R is hereditary.
(ii) Every free R-module is hereditary.
(iii) Every projective R-module is hereditary.
(iv) Every quotient of an injective R-module is injective.
Proof. (i) =⇒ (ii) is immediate from Corollary 3.56.
(ii) =⇒ (iii): Suppose that every free R-module is hereditary. Then if P is a
projective R-module, P is a submodule of a free module, hence a submodule of a
hereditary module, hence itself hereditary.
(iii) =⇒ (i): R is a projective R-module.
(iv) ⇐⇒ (iii): Let P 0 be a submodule of a projective R-module P ; call the
inclusion j. We will use Lemma 3.57a) : let I be an injective module, q : I → I 0 a
surjection, and f : P 0 → I 0 a module map. By assumption I 0 is injective, so there
is h : P → I 0 such that h ◦ j = f . Since P is projective, there is k : P → I such
that q ◦ k = h. Then F = k ◦ j : P 0 → I lifts f : q ◦ F = q ◦ k ◦ j = h ◦ j = f .
(iii) =⇒ (iv): Using Lemma 3.57b) we may dualize the proof of (iv) =⇒ (iii). 
Theorem 3.59. Let R be a a PID and F a free R-module. Then any submodule
M of F is again free, of rank less than or equal to the rank of F . In particular R
is a hereditary ring.
L
Proof. Let N be a submodule of F = L i∈I R. By Theorem 3.55b) for all
i ∈ I there is an ideal Ji of R such that N ∼
= i∈I Ji . Since R is a PID, each Ji is
is a free R-module of rank 0 or 1, so N is free of rank at most #I. 
Corollary 3.60. A projective module over a PID is free.
We expect that the following result is familiar to the reader as a special case of the
classification of (all) finitely generated modules over a PID, but while we are here
we may as well give a commutative algebraic proof.
Proposition 3.61. A finitely generated torsionfree module over a PID is free.
Proof. Let M be finitely generated and torsionfree. Certainly we may assume
that M is nonzero. Let X be a finite generating set for M with 0 ∈ / X. Let S ⊂ X
be a maximal R-linearly independent subset. Since M is torsionfree, S is not empty.
Let N = hSiR be the R-module spanned by S. Clearly N is free with basis S, and
we will be done if we can show N = M .
There is an annoying technicality here: we must check that S is finite. In fact,
let n = #X, and let S be any finite R-linearly independent subset and let s = #S.
We claim that s ≤ n. To see this, we tensor with the fraction field K of R, getting
ιK
Ks ∼= S ⊗R K → M ⊗R K ∼ = K m with m ≤ n. We may conclude that s ≤ m ≤ n
provided we know that ιK is injective. And this holds because K is a flat R-module,
which we will prove later on as a special case of the flatness of localization maps.
Until then, the reader must take this part of the proof on faith.
Say S = {x1 , . . . , xk }. Of course we are done already if S = X, so assume that
X \ S is nonempty. For each y ∈ X \ S there exist ry , r1 , . . . , rk ∈ R, not all zero,
such that ry y = r1 x1 + . . . + rk xk . Then ry 6= 0, since otherwise by the linear
independence of S all the ri would be zero. In other words, we have Q shown that
for each y ∈ X \ S there exists ry ∈ R• such that ry y ∈ N . Put r := y∈X\S ry .
9. ORDINAL FILTRATIONS AND APPLICATIONS 69

Then rX ⊂ N and thus rM ⊂ N . Now consider the R-module homomorphism


L : M → M given by multiplication by r: x 7→ rx. We have just established that
L(M ) ⊂ N , so we may regard L as a homomorphism M → N . Moreover, since M
is torsionfree, L is injective, and therefore L realizes M as a submodule of the free
R-module N . By Theorem 3.59 we conclude that M is free. 
Exercise 3.68. Let R be a ring with the following property: every submodule
of a finitely generated free R-module is free. Show that R is a principal ring (i.e.,
every ideal of R is principal).
An R-module M is semihereditary if every finitely generated R-submodule N of
M is projective. Thus a Noetherian semihereditary module is hereditary. A ring R
is semihereditary if R is a semihereditary R-module, or equivalently every finitely
generated ideal of R is projective as an R-module.
Example 3.62. A domain in which every finitely generated ideal is principal
is semihereditary. These are called Bézout domains and will be studied later on.
Theorem 3.63. Let {Mi }i∈I a familyL of semihereditary R-modules, and let N
be a finitely generated R-submodule ofL i∈I Mi . Then for all i ∈ I there is an
R-submodule Hi of Mi such that N ∼ = i∈I Hi .
Proof. The proof of Theorem 3.55a) goes through verbatim. 
Theorem 3.64. Let R be a domain in which every finitely generated ideal is
principal, and let F be a free R-module. Then any finitely generated submodule N
of F is free, of rank less than or equal to the rank of F .
Proof. One can adapt the proof of Theorem 3.60, using Theorem 3.63 in place
of Theorem 3.55. 
Theorem 3.65. Let R be a domain in which every finitely generated ideal is
principal. Then every finitely generated torsionfree R-module is principal.
Proof. The argument is the same as that of Proposition 3.61 (a special case),
using Theorem 3.64 in place of Theorem 3.59. 
Theorem 3.66. (F. Albrecht) Let R be a semihereditary ring, F a free R-
module, and P a finitely generated submodule of F .
a) P is isomorphic to a finite direct sum of finitely generated ideals of R.
b) In particular, P is a finitely generated projective module.
c) If R is a domain with fraction field K and F is free of finite rank n, then the
rank of P – i.e., dimK P ⊗R K – is at most n.
Exercise 3.69. Use Theorem 3.63 to prove Theorem 3.66.
9.3. Big modules.
Lemma 3.67. (Kaplansky) Let R be a ring, and let F L be an R-module which
is a direct sum of countably generated submodules: say F = λ∈Λ Eλ . Then every
direct summand of F is again a direct sum of countably generated submodules.
Proof. We claim that there is an ordinal filtration {Fi }i≤α on F satisfying
all of the following properties. (i) For all i < α, Fi+1 /Fi is countably generated.
(ii) If Mi = Fi ∩ M, Ni = Fi ∩ N , then Fi = Mi ⊕ Ni . L
(iii) For each i there is a subset Λi of Λ such that Fi = λ∈Λi Λi .
70 3. MODULES

sufficiency of claim: If so, {Mi }i≤α is an ordinal filtration on M . Moreover,


since Mi ⊂ Mi+1 are both direct summands of F , Mi is a direct summand of Mi+1 .
The Transfinite Dévissage Lemma (Lemma 3.52) applies to give

M∼
M
= Gr(M ) = Mi+1 /Mi .
i<α

Moreover, for all i < α we have


Fi+1 /Fi = (Mi+1 ⊕ Ni+1 )/(Mi ⊕ Ni ) ∼
= Mi+1 /Mi ⊕ Ni+1 /Ni ,
which shows that each successive quotient Mi+1 /Mi is countably generated. There-
fore M is a direct sum of countably generated submodules.
proof of claim: We will construct the filtration by transfinite induction. The
base case and the limit ordinal induction step are forced upon us by the definition
of ordinal filtration: we must have F0 = {0}, and for any limit S ordinal β ≤ α,
assuming we have defined Fi for all i < β we must have Fβ = i<β Fi .
So consider the case of a successor ordinal β = β 0 + 1. Let Q1 be any Eλ
which is not contained in Fβ 0 . (Otherwise we have Fβ 0 = F and we may just define
Fi = F for all β ≤ i ≤ α.) Let x11 , x12 , . . . be a sequence of generators of Q1 , and
decompose x11 into its M - and N -components. Let Q2 be the direct sum of the
finitely many Eλ which are necessary to write both of these components, and let
x21 , x22 , . . . be a sequence of generators for Q2 . Similarly decompose x12 into M
and N components, and let Q3 be the direct sum of the finitely many Eλ needed
to write out these components, and let x31 , x32 , . . . be a sequence of generators
of Q3 . We continue to carry out this procedure for all xij , proceeding accord-
ing to a diagonal enumeration of Z+ × Z+ : i.e., x11 , x12 , x21 , x13 , x22 , x31 , . . .. Put
Fβ = hFβ 0 , {xij }i,j∈Z+ iR . This works! 

For a cardinal number κ, we say that a module is κ-generated if it admits a


generating set of cardinality at most κ.
Exercise 3.70. (Warfield) Let κ be an infinite cardinal. Formulate and prove a
version of Lemma 3.67 in which “countably generated” is replaced by “κ-generated”.
Theorem 3.68. (Kaplansky) For a ring R, let Pc be the class of countably
generated projective R-modules. For an R-module M , the following are equivalent:
(i) M admits an ordinal filtration of class Pc .
(ii) M is a direct sum of countably generated projective submodules.
(iii) M is projective.
Proof. (i) ⇐⇒ (ii) follows immediately from Lemma 3.52.
(ii) =⇒ (iii): any direct sum of projective modules is projective.
(iii) =⇒ (ii): If M is projective, let F be a free R-module with F = M ⊕
N . Certainly F is a direct sum of countably generated submodules (indeed, of
singly generated submodules!), so by Lemma 3.67 M is a direct sum of a family of
countably generated submodules, each of which must be projective. 

While pondering the significance of this result, one naturally inquires:


Question 2. Is there a ring R and an R-module M which is not a direct sum
of countably generated submodules?
9. ORDINAL FILTRATIONS AND APPLICATIONS 71

Theorem 3.69. (Cohen-Kaplansky [CK51], Griffith) For a ring R, the follow-


ing are equivalent:
(i) Every R-module is a direct sum of cyclic (i.e., singly generated) R-modules.
(ii) Every R-module is a direct sum of finitely generated submodules.
(iii) R is an Artinian principal ideal ring.

Building on these results as well as work of Faith and Walker [FW67], R.B. Warfield
Jr. proved the following striking results.

Theorem 3.70. (Warfield [Wa]) Let R be a Noetherian ring which is not a


principal Artinian ring. Then for any cardinal κ, there exists a module M with the
following properties:
(i) M is not κ-generated, and
(ii) Any decomposition of M into the direct sum of nonzero submodules has only
finitely many direct summands.

The hypotheses of Theorem 3.70 apply for instance to the ring Z of integers and
yields, in particular, for any infinite cardinal κ a commutative group M which is
not a direct sum of κ-generated submodules.

Theorem 3.71. (Warfield [Wa]) For a ring R, the following are equivalent:
(i) Every R-module is a direct sum of cyclic submodules.
(ii) There exists a cardinal number κ such that every R-module is a direct sum of
κ-generated submodules.
(iii) R is a principal Artinian ring.

It is natural to wonder whether Theorem 3.70 can be strengthened in the following


way: an R-module M is indecomposable if it cannot be expressed as a direct sum
of two nonzero submodules.

Question 3. For which rings R do there exist indecomposable R-modules of


all infinite cardinalities?

However, Question 3 has turned out to be bound up with sophisticated set-theoretic


considerations. Namely, in a 1959 paper [Fu59], L. Fuchs claimed that there exist
indecomposable commutative groups of all infinite cardinalties, thus giving an affir-
mative answer to Question 3 for the ring R = Z. However, it was later observed (by
A.L.S. Corner) that Fuchs’ argument is valid only for cardinals κ less than the first
inaccessible cardinal. Exactly what an inaccessible cardinal is we do not wish to
say, but we mention that the nonexistence of inaccessible cardinals is equiconsistent
with the standard ZFC axioms of set theory (in other words, if the ZFC axioms
are themselves consistent, then ZFC plus the additional axiom that there are no
inaccessible cardinals remains consistent) but that nevertheless set theorists have
reasons to believe in them. See also [Fu74] in which these issues are addressed
and he proves that there is an indecomposable commutative group of any infinite
nonmeasurable cardinality (note: accessible implies nonmeasurable).

Question 4. Is there a ring R and a projective R-module M which is not a


direct sum of finitely generated submodules?
72 3. MODULES

Again the answer is yes. A very elegant example was given by Kaplansky (unpub-
lished, apparently).11 Namely that R be the ring of all real-valued continuous func-
tions on the unit interval [0, 1], and let I be the ideal of functions f : [0, 1] → R which
vanish near zero: i.e., for which there exists  = (f ) > 0 such that f |[0,(f )] = 0.
Exercise 3.71. Show the ideal I defined above gives a projective R-module
which is not the direct sum of finitely generated submodules. (Suggestions: (i) to
show that I is projective, use the Dual Basis Lemma. (ii) A slick proof of the fact
that I is not a direct sum of finitely generated submodules can be given by Swan’s
Theorem using the contractibility of the unit interval.)
Lemma 3.72. Let M be a projective module over the local ring R, and let x ∈ M .
There is a direct summand M 0 of M such that M 0 contains x and M 0 is free.
Proof. Let F be a free module with F = M ⊕ N . Choose a basis B = {ui }i∈I
of F with respect to which the element x of M has the minimal possible number
of nonzero coordinates. Write
x = r1 u1 + . . . + rn un , ri ∈ R• .
P Pn−1
Then for all 1 ≤ i ≤ n, ri ∈ / j6=i Rrj . Indeed, if say rn = i=1 si ri , then
Pn−1
x = i=1 ri (ui + si un ), contradicting the minimality of the chosen basis.
Now write ui = yi + zi with yi ∈ M, zi ∈ N , so
X X
(6) x= ri ui = ri yi .
i i
We may write
n
X
(7) yi = cij uj + ti ,
j=1

with ti a linear combination of elements of B \ {u1 , . . . , un }. Substituting (7) into


(6) and projecting onto M gives the relations
n
X
ri = cji rj ,
j=1

or equivalently, for all i, X


(1 − cii )ri = cji rj .
j6=i
If for any i and j, then one of the coefficients of rj in the above equation is a unit
of R, then dividing through by it expresses rj as an R-linear combination of the
other ri ’s, which as above is impossible. Therefore, since R is local, each coefficient
must lie in the maximal ideal of R:
∀i, 1 − cii ∈ m, ∀i 6= j, cij ∈ m.
It follows that the determinant of the matrix C = (cij ) is congruent to 1 modulo m,
hence invertible: if x ∈ m and 1 + x is not invertible, then 1 + x = y for y ∈ m, so
1 = y − x ∈ m, contradiction. Therefore replacing u1 , . . . , un in B with y1 , . . . , yn
still yields a basis of F . It follows that M 0 = hy1 , . . . , yn iR is a direct summand of
F hence also of M which is a free module containing x. 
11Warm thanks to Gjergji Zaimi for bringing this important example to my attention.
10. TOR AND EXT 73

Theorem 3.73. (Kaplansky) Let (R, m) be a local ring, and let P be any
projective R-module. Then P is free.
Proof. Step 1: Since by Theorem 3.68 P is a direct sum of countably gen-
erated projective submodules, we may as well assume that P itself is countably
generated.
Step 2: Suppose M = h{xn }∞ n=1 iR is a countably generated projective module over
the local ring R. By Lemma 3.72, M = F1 ⊕ M1 with F1 free containing x1 . Note
that M1 is again projective and is generated by the images {x0n }∞
n=2 of the elements
xn under the natural projection map M → M1 . So reasoning as above, we may
write M2 = F2 ⊕ M2 with F2 free containing x02 . Continuing in this manner, we get

M
M= Fn ,
n=1

so M is free. 

Exercise 3.72. Give an example of a (necessarily infinitely generated) module


over a local PID which is flat but not free.

10. Tor and Ext


10.1. Co/chain complexes.

Let R be a ring. A chain complex C• of R-modules is a family {Cn }n∈Z of


R-modules together with for all n ∈ Z, an R-module map dn : Cn → Cn−1 such
that for all n, dn−1 ◦ dn = 0. (It is often the case that Cn = 0 for all n < 0, but
this is not a required part of the definition.)
An example of a chain complex of R-modules is any long exact sequence. How-
ever, from the perspective of homology theory this is a trivial example in the follow-
ing precise sense: for any chain complex we may define its homology modules:
for all n ∈ Z, we put
Hn (C) = Ker(dn )/ Im(dn+1 ).
Example 3.74. Let X be any topological space. For any ring R, we have the
singular chain complex S(X)• : S(X)n = 0 for n < 0, and for n ≥ 0, S(X)n is
the free R-module with basis the set of all continuous maps ∆n → X, where ∆n is
the standard n-dimensional simplex. A certain carefully defined alternating sum of
restrictions to faces of ∆n gives rise to a boundary map dn : S(X)n → S(X)n−1 ,
and the indeed the homology groups of this complex are nothing else than the singular
homology groups Hn (X, R) with coefficients in R.
If C• and D• are two chain complexes of R-modules, a homomorphism η : C• →
D• is given by maps ηn : Cn → Dn for all n rendering the following infinite ladder
commutative:
INSERT ME!.
In this way one has evident notions of a monomorphism and epimorphisms of
chain complexes. In fact the chain complexes of R-modules form an abelian cate-
gory and thus these notions have a general categorical meaning, but it turns out
they are equivalent to the much more concrete naive conditions: η is a monomor-
phism iff each ηn is injective and is an epiomorphism iff each ηn is surjective.
74 3. MODULES

In particular it makes sense to consider a short exact sequence of chain com-


plexes:
0 −→ A• −→ B• −→ C• .
Here is the first basic theorem of homological algebra.
Theorem 3.75. Let
f g
0 −→ A• −→ B• −→ C• −→ 0
be a short exact sequence of chain complexes of R-modules. Then for all n ∈ Z
there is a natural connecting homomorphism ∂ : Hn (C) → Hn−1 (A) such that
g ∂ f g ∂ f
. . . −→ Hn+1 (C) −→ Hn (A) → Hn (B) −→ Hn (C) −→ Hn−1 (A) −→ . . .
is exact.
Proof. No way. See [W, Thm. 1.3.1]. 
Moreover, the homology modules Hn are functors: if f : C• → D• is a morphism
of chain complexes, there are induced maps on the homology groups
Hn (f ) : Hn (C) → Hn (D).
Example 3.76. Let f : X → Y be a continuous map of topological spaces.
Then for any basic n-chain ∆n → X in S(X)n , composition with f gives a basic
n-chain ∆n → Y in S(Y )n and thus a homomorphism of chain complexes S(f ) :
S(X)• → S(Y )• . There are induced maps on homology, namely the usual maps
Hn (f ) : Hn (X, R) → Hn (Y, R).
There is an entirely parallel story for cochain complexes of R-modules, which are
exactly the same as chain complexes but with a different indexing convention: a
cochain complex C • consists of for each n ∈ Z+ an R-module C n and a “coboundary
map” dn : C n → C n+1 . To any cochain complex we get cohomology modules:
for all n ∈ Z, put
H n (C) = Ker(dn )/ Im(dn−1 ).
The rest of the discussion proceeds in parallel to that of chain complexes (including
the realization of singular cohomology as a special case of this construction).
10.2. Chain homotopies.

Let C• , D• be two chain complexes, and let f, g : C• → D• be two homomor-


phisms between them. We say that f and g are chain homotopic if there exist
for all n ∈ Z+ R-module maps sn : Cn → Dn+1 such that
fn − gn = dn+1 sn + sn−1 dn .
The sequence {sn } is called a chain homotopy from f to g.
Exercise 3.73. Show: chain homotopy is an equivalence relation on morphisms
from C• to D• .
What on earth is going on here? Again topology is a good motivating example: we
say that two maps f, g : X → Y are homotopic if there exists a continuous map
F : X × [0, 1] → Y such that for all x ∈ X, F (x, 0) = f (x) and F (x, 1) = g(x).
This is an equivalence relation and is generally denoted by f ∼ g. We then define
10. TOR AND EXT 75

two topological spaces to be homotopy equivalent if there exist maps ϕ : X → Y


and ψ : Y → X such that
ψ ◦ ϕ ∼ 1X , ϕ ◦ ψ ∼ 1Y .
(We say that ϕ : X → Y is a homotopy equivalence if there exists a map ψ as
above.) E.g. a space is contractible if it is homotopy equivalent to a single point.
One of the basic tenets of algebraic topology is that it aspires to study topo-
logical spaces only up to homotopy equivalence. That is, all of the fundamen-
tal invariants of spaces should be the same on homotopy equivalent spaces and
homomorphisms between these invariants induced by homotopic maps should be
identical. Especially, if f : X → Y is a homotopy equivalence, the induced maps
Hn (f ) : Hn (X) → Hn (Y ) should be isomorphisms. In fact, if f, g : X → Y are ho-
motopic, the induced morphisms S(f ), S(g) : S(X)• → S(Y )• are chain homotopic.
So the following result ensures that the induced maps on homology are equal.
Proposition 3.77. If f, g : C• → D• are chain homotopic, then for all n ∈ Z,
Hn (f ) = Hn (g).
Proof. Replacing f and g by f − g and 0, it is enough to assume that there
exists a chain homotopy s from f to the zero map – i.e., for all n fn = dn+1 sn +
sn−1 dn – and show that f induces the zero map on homology. So take x ∈ Hn (C).
Then x is represented by an element of Cn lying in the kernel of dn , so
fn (x) = dn+1 sn x + sn−1 dn x = dn+1 sn x + 0 = dn+1 sn x.
Thus fn (x) lies in the image of dn+1 Dn+1 → Dn so represents 0 ∈ Hn (D). 

10.3. Resolutions.

Let M be an R-module. A left resolution of M is an infinite sequence {Ai }∞


i=0
of R-modules, for all n ∈ N an R-module map An+1 → An and an R-module map
A0 → M such that the sequence
. . . −→ An+1 → An −→ . . . −→ A1 −→ A0 −→ M −→ 0
is exact. By abuse of notation, we often speak of “the resolution A• . Dually, a
right resolution of M is an infinite sequence {B i }∞
i=0 of R-modules, for all n ∈ N
an R-module map B n → B n+1 and an R-module map M → B 0 such that the
sequence
0 −→ M −→ B 0 −→ B 1 −→ . . . −→ B n −→ B n1 . . .
is exact. We speak of “the resolution B • ”.

A projective resolution of M is a left resolution A• such that each An is projec-


tive. A injective resolution of M is a right resolution B • such that each B n is
injective. (Exactly why we are not interested in left injective resolutions and right
projective resolutions will shortly become clear.)
Theorem 3.78. (Existence of resolutions) Let M be an R-module.
a) Since every R-module is the quotient of a projective (indeed, of a free) module,
M admits a projective resolution.
b) Since every R-module can be embedded in an injective module, M admits an
injective resolution.
76 3. MODULES

Proof. a) Choose a projective module P0 , a surjection 0 : P0 → M , and


put M0 = ker(0 ). Inductively, given Mn−1 , we choose a projective module Pn , a
surjection n : Pn → Mn−1 , and put Mn = ker(0 ). As our map dn : Pn → Pn−1
we take the composite

n ker(n−1 )
Pn −→ Mn−1 −→ Pn−1 .
We claim that the resulting sequence
. . . −→ Pn+1 → Pn −→ . . . −→ P1 −→ P0 −→ M −→ 0
is exact. It is certainly exact at M . If x ∈ P0 and 0 (x) = 0, then x0 ∈ M0 .
Lifting x0 via the surjection 1 to x1 ∈ P1 , we find d1 (x1 ) = 1 (x1 ) = x0 , so
ker(0 ) ⊂ Im(d1 ). Conversely, since d1 factors through ker(0 ), it is clear that
Im(d1 ) ⊂ ker(0 ). Exactly the same argument verifies exactness at Pn for each
n > 0, so P• is a projective resolution of M .
b) We leave the proof of this part to the reader as an exercise, with the following
comforting remark: the notion of an injective module is obtained from the notion
of a projective module by “reversing all the arrows”, which is the same relationship
that a left resolution bears to a right resolution. Therefore it should be possible to
prove part b) simply by holding up the proof of part a) to a mirror. (And it is.) 

Theorem 3.79. (Comparison theorem for resolutions)


a) Let P• be a projective resolution of the R-module M . Let N be another R-module
and f−1 : M → N be an R-module map. Then for every left resolution A• of N
there exists a homomorphism η from the chain complex P• → M → 0 to the chain
complex A• → N → 0. Moreover η is unique up to chain homotopy.
b) Let E • be an injective resolution of the R-module N . Let M be another R-module
and f 0 : M → N be an R-module map. Then for every right resolution A• of M
there exists a homomorphism η from the chain complex 0 → M → A• to the chain
complex 0 → N → E • . Moreover η is unique up to chain homotopy.
Proof. No way. See [W, Thms. 2.2.6 and 2.3.7]. 

Exercise 3.74. Let F be a covariant additive functor on the category of R-


modules. Let C• and D• be two chain complexes of R-modules and f, g : C• → D•
be two homomorphisms between them.
a) Show that F C• and F D• are chain complexes and there are induced chain ho-
momorphisms F f, F g : F C• → F D• .
b) Show that if f and g are chain homotopic, so are F f and F g.
(Suggestion: Show that it makes sense to apply F to a chain homotopy s.)
10.4. Derived functors.

Let us consider covariant, additive functors F from the category of R-modules to it-
self. (Recall that additive means that for any M, N , the induced map Hom(M, N ) →
Hom(F (M ), F (N )) is a homomorphism of commutative groups.)
Exercise 3.75. For any additive functor F and any chain complex C• of R-
modules, F C• is again a chain complex.
(Hint: an additive functor takes the zero homomorphism to the zero homomor-
phism.)
10. TOR AND EXT 77

Thus if
0 −→ M1 −→ M2 −→ M3 −→ 0
is a short exact sequence of R-modules, then
0 −→ F (M1 ) −→ F (M2 ) −→ F (M3 ) −→ 0
is necessarily a complex of modules but not necessarily exact: it may have nonzero
homology.
Example 3.80. For any ring R, the functor F (M ) = M ⊕ M is exact. For
R = Z the functor F (M ) = M ⊗ Z/2Z is not exact: for instance it takes the short
exact sequence
·2
0 −→ Z −→ Z −→ Z/2Z −→ 0
to the complex
·2
0 −→ Z/2Z −→ Z/2Z → Z/2Z −→ 0,
but multiplication by 2 on Z/2Z is not an injection.
Although an exact functor is a thing of beauty and usefulness to all, it turns out
that from a homological algebraic point of view, it is the functors which are “half
exact” which are more interesting: they give rise to co/homology theories.

An additive functor F is right exact if for any exact sequence of the form
M1 −→ M2 −→ M3 −→ 0,
the induced sequence
F M1 −→ F M2 −→ F M3 −→ 0
is again exact. This much was true for the functor F (M ) = M ⊗ Z/2Z, at least for
the sequence we chose above. In fact this holds for all tensor products.
Proposition 3.81. For any ring R and any R-module N , the functor F (M ) =
M ⊗R N is right exact.
Exercise 3.76. Prove Proposition 3.81
We have also the dual notion of an additive functor F being left exact: for any
exact sequence of the form
0 → M1 → M 2 → M3 ,
the induced sequence
0 → F M1 → F M2 → F M3
is again exact.

We now wish to press our luck a bit by extending this definition to contravari-
ant functors. Here a little abstraction actually makes me less confused, so I will
pass it along to you: we say that a contravariant functor F from the abelian cate-
gory C to the abelian category D is left exact (resp. right exact) if the associated
covariant functor F opp : C opp → D is left exact (resp. right exact). Concretely, a
contravariant functor F from R-modules to R-modules is left exact if every exact
sequence of the form
M1 → M2 → M 3 → 0
78 3. MODULES

is transformed to an exact sequence


0 → F M3 → F M2 → F M1 .
(And similarly for right exact contravariant functors.)
Proposition 3.82. Let R be a ring and X be an R-module.
a) The functor M 7→ Hom(X, M ) is covariant and left exact.
(Recall that it is exact iff X if projective.)
b) The functor M 7→ Hom(M, X) is contravariant and left exact.
(Recall that it is exact iff X is injective.)
Exercise 3.77. Prove Proposition 3.82.
Let F be a right exact additive functor on the category of R-modules. We will
define a sequence {Ln F }n∈N of functors, with L0 F = F , called the left derived
functors of F . The idea here is that the left-derived functors quantify the failure
of F to be exact.

Let M be an R-module. We define all the functors Ln M at once, as follows:


first we choose any projective resolution P• → M → 0 of M . Second we take away
the M , getting a complex P• which is exact except at P0 , i.e.,
H0 (P ) = P0 / Im(P1 → P0 ) = P0 / Ker(P0 → M ) = M,
∀n > 0, Hn (P ) = 0.
Third we apply the functor F getting a new complex F P• . And finally, we take
homology of this new complex, defining
(Ln F )(M ) := Hn (F P• ).
Now there is (exactly?) one thing which is relatively clear at this point.
Proposition 3.83. We have (L0 F )(M ) = F M .
Proof. Since P1 → P0 → M → 0 is exact and F is right exact, F P1 → F P0 →
F M → 0 is exact, hence
Im(F P1 → F P0 ) = Ker(F P0 → F M ).
Thus
(L0 F )(M ) = H0 (F P• ) = Ker(F P0 → 0)/ Im(F P1 → F P0 )
= F P0 / Ker(F P0 → F M ) = F M.

Before saying anything else about the left derived functors Ln F , there is an obvious
point to be addressed: how do we know they are well-defined? On the face of it,
they seem to depend upon the chosen projective resolution P• of M , which is very
far from being unique. To address this point we need to bring in the Comparison
Theorem for Resolutions (Theorem 3.79). Namely, let P•0 → M → 0 be any other
projective resolution of M . By Theorem 3.79, there exists a homomorphism of
chain complexes η : P• → P•0 which is unique up to chain homotopy. Interchanging
the roles of P•0 and P• , we get a homomorphism η 0 : P•0 → P• . Moreover, the
composition η 0 ◦ η is a homomorphism from P• to itself, so by the uniqueness η 0 ◦ η
is chain homotopic to the identity map on P• . Similarly η ◦ η 0 is chain homotopic to
the identity map on P•0 , so that η is a chain homotopy equivalence. By Exercise 3.71,
10. TOR AND EXT 79

F η : F P• → F P•0 is a chain homotopy equivalence, and therefore the induced maps


on homology Hn (F η) : Hn (F P• ) → Hn (F P•0 ) are isomorphisms. Thus we have
shown that two different choices of projective resolutions for M lead to canonically
isomorphic modules (Ln F )(M ).
Exercise 3.78. Suppose M is projective. Show: for any right exact functor F
and all n > 0, (Ln F )(M ) = 0.
The next important result shows that a short exact sequence of R-modules induces
a long exact sequence involving the left-derived functors and certain connecting
homomorphisms (which we have not defined and will not define here).
Theorem 3.84. Let
(8) 0 −→ M1 −→ M2 −→ M3 −→ 0
be a short exact sequence of R-modules, and let F be any left exact functor on the
category of R-modules. Then:
a) There is a long exact sequence
(9)
∂ ∂
. . . → (L2 F )(M3 ) → (L1 F )(M1 ) → (L1 F )(M2 ) → (L1 F )(M3 ) → F M1 → F M2 → F M3 → 0.
b) The above construction is functorial in the following sense: if 0 −→ N1 −→
N2 −→ N3 −→ 0 is another short exact sequence of R-modules and we have maps
Mi → Ni making a “short commutative ladder”, then there is an induced “long
commutative latter” with top row the long exact sequence associated to the first
short exact sequence and the bottom row the long exact sequence associated to the
second short exact sequence.
Proof. No way. See [W, Thm. 2.4.6]. 

Remark: One says that (9) is the long exact homology sequence associated to
the short exact sequence (8).

Now, dually, if F is a right exact functor on the category of R-modules, we may


define right derived functors Rn F . Namely, for an R-module M , first choose
an injective resolution 0 → M → E • , then take M away to get a cochain com-
plex E • , then apply F to get a cochain complex F E • , and then finally define
(Rn F )(M ) = H n (F E • ). In this case, a short exact sequence of modules (8) in-
duces a long exact cohomology sequence
(10)
∂ ∂
0 → F M1 → F M2 → F M3 → (R1 F )(M1 ) → (R1 F )(M2 ) → (R1 F )(M3 ) → (R2 F )(M1 ) . . .
Exercise 3.79. Suppose M is injective. Show: for any left exact functor F
and all n > 0, we have (Rn F )(M ) = 0.
10.5. Tor.

Let M, N be R-modules, and let F : N → M ⊗R N be the functor “tensor with


M ”. By X.X F is right exact so has left derived functors (Ln F ). By definition, for
all n ∈ N,
Torn (M, N ) := (Ln F )(N ).
80 3. MODULES

Now un/fortunately the situation is even a little richer than the general case of
left-derived functors discussed above. Namely, the tensor product is really a bi-
functor: i.e., a functor in M as well as in N , additive and covariant in each variable
separately. So suppose we took the right-derived functors of M 7→ M ⊗R N and
applied them to M : this would give us Torn (N, M ). So it is natural to ask: how
does Torn (M, N ) compare to Torn (N, M )? Since for n = 0 we have that M ⊗R N
is canonically isomorphic to N ⊗R M , it is natural to hope that the Tor functors
are symmetric. And indeed this turns out to be the case.
Theorem 3.85. (Balancing Tor) For any R-modules M and N and all n ≥ 0,
there are natural isomorphisms Torn (M, N ) = Torn (N, M ).
Proof. No way. See [W, Thm. 2.7.2]. 

Exercise 3.80. In order to use the Universal Coefficient Theorem (for homol-
ogy) in algebraic topology, one needs to know the values of Tor1 (M, N ) for any two
finitely generated Z-modules M and N .
a) Show that for any m, n ∈ Z+ , Tor1 (Z/mZ, Z/nZ) ∼ = Z/ gcd(m, n)Z.
b) Show that for all Z-modules N , Tor1 (Z, N ) = 0.
c) Explain how the structure theorem for finitely generated Z-modules reduces the
problem of computation of Tor1 (M, N ) for any finitely generated M and N to the
two special cases done in parts a) and b).
Exercise 3.81. Show: the Tor functors commute with direct limits in the sense
that for all n ∈ N, any directed system {Mi }i∈I of R-modules M and any R-module
N we have a canonical isomorphism
Torn (lim Mi , N ) → lim Torn (Mi , N ).
−→ −→
i i

(Suggestion: the case n = 0 is Proposition 3.9. Use this to show the general case
by brute force: i.e., take a projective resolution of N and track these isomorphisms
through the definition of Torn .)
10.6. Ext.

Let M, N be R-modules, and let F : N → Hom(M, N ). By Proposition X.X,


F is covariant and left exact. By definition, for all n ∈ N,
Extn (M, N ) = (Rn F )(N ).
But again, we have an embarrassment of riches: why didn’t we define the Ext
functors as the right-derived functors of the contravariant left exact functor G :
N → Hom(N, M )? Again, we can do this.
Theorem 3.86. (Balancing Ext) Let M and N be R-modules. Define functors
FM : N → Hom(M, N ) and GN : M → Hom(M, N ). Then for all n ≥ 0,
(Rn FM )(N ) = (Rn GN )(M ).
Proof. No way. See [W, Thm. 2.7.6]. 

Exercise 3.82. In order to use the Universal Coefficient Theorem (for coho-
mology) in algebraic topology, one needs to know the values of Ext1 (M, N ) for any
two finitely generated Z-modules M and N . Compute them.
11. MORE ON FLAT MODULES 81

Theorem 3.87. a) For an R-module P , the following are equivalent:


(i) P is projective.
(ii) Ext1R (P, B) = 0 for all R-modules B.
b) For an R-module E, the following are equivalent:
(i) E is injective.
(ii) Ext1R (A, E) = 0 for all R-modules A.
Theorem 3.88. a) For a ring R, the following are equivalent:
(i) R is hereditary.
(ii) Every R-module M admits a projective resolution of the form 0 → P1 → P0 →
M → 0.
(iii) For all R-modules M and N and all n ≥ 2, ExtnR (M, N ) = 0.
b) The conditions of part a) imply:
(iv) For all R-modules M and N and all n ≥ 2, TornR (M, N ) = 0.
c) If R is Noetherian, then (iv) =⇒ (i) and thus all are equivalent.
Proof. CITE. 
Exercise 3.83. Use Corollary 3.59 to show (i) =⇒ (ii) =⇒ (iii).
Theorem 3.89. For R-modules A and C, the following are equivalent:
(i) Every short exact sequence 0 → A → B →→ C → 0 splits.
(ii) Ext1R (C, A) = 0.
Proof. See e.g. [Ro, Thm. 7.31]. 

11. More on flat modules


Theorem 3.90. (Tensorial Criterion for Flatness) For an R-module M , the
following are equivalent:
(i) M is flat.
(ii) For every finitely generated ideal I of R the canonical map I ⊗R M → IM is
an isomorphism.
Proof. Fist note that the canonical map I ⊗R M → IM is always a surjection.
(i) =⇒ (ii): if M is flat, then since I ,→ R, I ⊗R M ,→ R ⊗R M = M , so

I ⊗R M → IM .
(ii) =⇒ (i): Every ideal of R is the direct limit of its finitely generated subideals,
so it follows from Proposition 3.9 and the exactness of direct limits that I ⊗M → M
is injective for all ideals I. Moreover, if N is an R-module and N 0 ⊂ N is an R-
submodule, then since N is the direct limit of submodule N 0 + F with F finitely
generated, to show that N 0 ⊗ M → N ⊗ M is injective we may assume
N = N 0 + hω1 , . . . , ωn iR .
We now proceed by dévissage: putting Ni = N 0 + hω1 , . . . , ωi iR , it is enough to
show injectivity at each step of the chain
N 0 ⊗ M → N1 ⊗ M → . . . → N ⊗ M,
and further simplifying, it is enough to show that if N = N 0 + Rω, then N 0 ⊗ M ,→
N ⊗ M . Let I be the “conductor ideal of N/N 0 ”, i.e., I = {x ∈ R | xω ∈ N 0 }, so
that we get a short exact sequence of R-modules
0 → N 0 → N → R/I → 0
82 3. MODULES

which gives rise to a long exact homology sequence


0
. . . → TorR
1 (M, R/I) → N ⊗ M → N ⊗ M → M/IM → 0.

Thus it suffices to prove TorR


1 (M, R/I) = 0. For this we consider the homology
sequence associated to
0 → I → R → R/I → 0,
namely
. . . → TorR R
1 (M, R) = 0 → Tor1 (M/R/I) → I ⊗ M → M → . . . ,

and from the injectivity of I ⊗ M → M we deduce TorR


1 (M, R/I) = 0. 
Theorem 3.91. (Homological Criterion for Flatness) For an R-module M ,
the following are equivalent:
(i) M is flat.
(ii) For every R-module N all i > 0, TorR i (M, N ) = 0.
(ii0 ) For every R-module N , TorR 1 (M, N ) = 0.
(iii) For every finitely generated ideal I of R, TorR1 (M, R/I) = 0.

Proof. (i) =⇒ (ii): This is a statement about projective resolutions, but


given that it is just about the most basic possible one. Namely, let L• → N → 0
be a projective resolution of N . Then
. . . → Ln ⊗ M → Ln−1 ⊗ M → . . . → L0 ⊗ M
is exact, so TorR
i (M, N ) = 0 for all i > 0.
(ii) =⇒ (ii0 ) and (ii0 ) =⇒ (iii) are both immediate.
(iii) =⇒ (i): For each finitely generated ideal I of R, the short exat sequence
0 → I → R → R/I → 0
of R-modules induces a long exact sequence in homology, which ends
. . . → TorR
1 (M, R/I) = 0 → I ⊗ M → M → M/IM → 0,

i.e., the map I ⊗ M → M is injective and thus induces an isomorphism I ⊗ M →
IM . Using the Tensorial Criterion for Flatness (Theorem 3.90), we conclude M is
flat. 
Corollary 3.92. (Direct limits preserve flatness) Let R be a ring and {Mi }i∈I
a directed system of flat R-modules. Then M = lim Mi is a flat R-module.
−→
Proof. For every R-module N , we have
TorR
1 (lim M , N) ∼
= TorR 1 (N, lim M ) = lim TorR ∼
1 (N, Mi ) = lim TorR
1 (Mi , N ) = −
lim 0 = 0.
−→ i −→ i −→ −→ →
Now apply the Homological Criterion for Flatness. 
Corollary 3.93. For a domain R, the following are equivalent:
(i) Every finitely generated torsionfree R-module is flat.
(ii) Every torsionfree R-module is flat.
Proof. Every submodule of a torsionfree R-module is torsionfree, and every
R-module is the direct limit of its finitely generated submodules. So the result
follows immediately from Proposition 3.92. 

Corollary 3.94. Let 0 → M 0 → M → M 00 → 0 be a short exact sequence of


R-modules, with M 00 flat. Then M 0 is flat iff M is flat.
11. MORE ON FLAT MODULES 83

Exercise 3.84. Use the Homological Criterion of Flatness to prove Corollary


3.94.
Exercise 3.85. In a short exact sequence of R-modules as in Corollary 3.94,
if M 0 and M are flat, must M 00 be flat?
Now recall that a finitely generated torsion free module over a PID is free (Propo-
sition 3.61). From this we deduce:
Corollary 3.95. A module over a PID is flat iff it is torsionfree.
Exercise 3.86. Let R be a domain and M a torsion R-module. Show that for
all R-modules N and all n ≥ 0, Torn (M, N ) is a torsion R-module.
Theorem 3.96. Let R be a PID and let M, N be R-modules.
a) For all n ≥ 2, Torn (M, N ) = 0.
b) Tor1 (M, N ) is a torsion R-module.
Proof. a) Choose a free module F0 and a surjection d0 : F0 → N . By X.X,
F1 = Ker(d0 ) is free, so we get a finite free resolution of N:
0 → F1 → F0 → N → 0.
Therefore we certainly have Torn (M, N ) for all M and all n ≥ 2.
b) Let {Mi }i∈I be the direct system of all finitely generated submodules of M . As
above, we have M = lim Mi , so
−→
Tor1 (M, N ) = Tor1 (lim Mi , N )) = lim Tor1 (Mi , N ).
−→ −→
By Corollary 3.95, each Mi which is torsionfree is flat, hence Tor1 (Mi , N ) = 0. Thus
the only possible contribution to lim Tor1 (Mi , N ) comes from torsion modules Mi ,
−→
and by Exercise 3.83, Mi torsion implies Tor1 (Mi , N ) torsion. Thus Tor1 (M, N ) is
a direct limit of torsion modules, hence itself a torsion module. 

Theorem 3.97. (Equational Criterion for Flatness) Let M be an R-module.


a) Suppose M is flat, and that we are given r, n ∈ Z+ , a matrix A = (aij ) ∈
Mr×n (R) and elements x1 , ldots, xn ∈ M such that
X
∀1 ≤ i ≤ r, aij xj = 0.
j

Then there are s ∈ Z+ , bjk ∈ R and yk ∈ M (for 1 ≤ j ≤ n and 1 ≤ k ≤ s) such


that
X
∀i, k, aij bj = 0
j

and
X
∀j, xj = bjk yk = 0.
j

Thus the solutions in a flat module of a system of linear equations with R-coefficients
can be expressed as a linear combination of solutions of the system in R.
b) Conversely, if the above conditions hold for a single equation (i.e., with r = 1),
then M is a flat R-module.
84 3. MODULES

Proof. a) Let ϕ : Rn → Rr be the linear map corresponding to multiplication


by the matrix A and let ϕM : M n → M r be the same for M , so that ϕM = ϕ ⊗ 1M .
Let K = Ker ϕ. Since M is flat, tensoring with M preserves exact sequences, thus
the sequence
ι⊗1 ϕ
K ⊗R M → M n → M r
is exact. By our hypothesis we have ϕM (x1 , . . . , xn ) = 0, so that we may write
s
!
X
(x1 , . . . , xn ) = (ι ⊗ 1) βk ⊗ y k
k=1
with βk ∈ K and yk ∈ M . Writing out each βk as an element (b1k , . . . , bnk ) ∈ Rn
gives the desired conclusion.
b) We will use the Tensorial Criterion for Flatness to show that M is flat. Let
I = ha1 , . . . , an i be a P
finitely generated ideal of R. We may Pnwrite an arbitrary
n
element z of I ⊗ M as i=1 ai ⊗ mi with mi ∈ M . Let z = i=1 ai mi denote the
imagePof z in IM ⊂ M . We want to show that z = 0 implies z = 0, so suppose
that
P i ai mi = 0. By hypothesis, there P exist bij ∈ R and yj ∈ M such that for all
j, i ai bij = 0 and for all i, mi = j bij yj . Thus
!
X XX X X X
z= ai ⊗ mi = ai bij ⊗ yj = ai bij ⊗ yj = 0 ⊗ yj = 0.
i i j j i j


As an application, we can now improve Theorem 3.51 by weakening the hypothesis
of “finite presentation” to the simpler one of “finite generation”.
Theorem 3.98. Let M be a finitely generated flat module over the local ring
(R, m). Then for all n ∈ Z+ , x1 , . . . , xn are elements of M such that the images in
R/m are R/m-linearly independent, then x1 , . . . , xn are R-linearly independent.
Proof. We go by induction on n. Suppose first that n = 1, in which case
it is sufficient to show that a1 ∈ R, a1 x1 6= 0 implies a1 = 0. By the Equational
P for Flatness, there exist b1 , . . . , bs ∈ R such that abi = 0 for all i and
Criterion
x1 ∈ i bi M . By assumption, x1 does not lie in mM , so that for some i we must
have bi ∈ R× , and then abi = 0 implies a = 0.
Now suppose n > 1, and let a1 , . . . , an ∈ R are such that a1 x1 + . . . + an xn = 0.
Again using the Equational P there are bij ∈ R and y1 , . . . , ys ∈
P Criterion for Flatness,
M such that for all j, ai bij = 0 and xi = bij yj . Since the set of generators
is minimal, by Nakayama’s Lemma their images in M/mM must be R/m-linearly
independent. In particular xn ∈ / mM , so that at least one bnj is a unit. It follows
Pn−1
that there exist c1 , . . . , cn−1 ∈ R such that an = i=1 ai ci . Therefore
a1 (x1 + c1 xn ) + . . . + an−1 (xn−1 + cn−1 xn ) = 0.
The images in M/mM of the n − 1 elements x1 + c1 xn , . . . , xn−1 + cn−1 xn are
R/m-linearly independent, so by induction a1 = . . . = an−1 = 0. Thus an = 0. 
Theorem 3.99. For a finitely generated module M over a local ring R, the
following are equivalent:
(i) M is free.
(ii) M is projective.
(iii) M is flat.
11. MORE ON FLAT MODULES 85

Proof. For any module over any ring we have (i) =⇒ (ii) =⇒ (iii). So
suppose that M is a finitely generated flat module over the local ring (R, m). Let
(x1 , . . . , xn ) be a set of R-module generators for M of minimal cardinality. By
Nakayama’s Lemma the images of x1 , . . . , xn in R/m are R/m-linearly independent,
and then Theorem 3.98 implies that x1 , . . . , xn is a basis for M as an R-module. 
A ring R is called absolutely flat if every R-module is flat.
Exercise 3.87. Show: a quotient of an absolutely flat ring is absolutely flat.
Proposition 3.100. For a ring R, the following are equivalent:
(i) R is absolutely flat.
(ii) For every principal ideal I of R, I 2 = I.
(iii) Every finitely generated ideal of R is a direct summand of R.
Proof. (i) =⇒ (ii): Assume R is absolutely flat, and let I = (x) be a
principal ideal. Tensoring the natural inclusion (x) → R with R/(x), we get a an
injection (x)⊗R R/(x) → R/(x). But this map sends x⊗r 7→ xr +(x) = (x), so it is
identically zero. Therefore its injectivity implies that 0 = (x) ⊗R R/(x) ∼
= (x)/(x2 ),
2
so (x) = (x ).
(ii) =⇒ (iii): Let x ∈ R. Then x = ax2 for some a ∈ R, so putting e = ax we
have e2 = a2 x2 = a(ax2 ) = ax = e, so e is idempotent, and (e) = (x). In general,
for any two idempotents e, f , we have he, f i = (e + f − ef ). Hence every finitely
generated ideal is principal, generated by an idempotent element, and thus a direct
summand.
(iii) =⇒ (i): Let M be an R-module, and let I be any finitely generated
ideal of R. By assumption, we may choose J such that R = I ⊕ J. Therefore J is
projective, so Tor1 (R/I, M ) = Tor1 (J, M ) = 0. By the Homological Criterion for
Flatness, M is flat. 
Exercise 3.88. Show: a (finite or infinite) product of absolutely flat rings is
absolutely flat.
The following striking result came relatively late in the game: it is due indepen-
dently to Govorov [Gov65] and Lazard [La64].
Theorem 3.101. (Govorov-Lazard) For a module M over a ring R, the fol-
lowing are equivalent:
(i) M is flat.
(ii) There exists a directed family {Fi }i∈I of finitely generated free submodules of
M such that M = lim Fi .
−→
Proof. (i) =⇒ (ii): Suppose M = lim Fi is a direct limit of free modules.
−→
Then in particular M is a direct limit of flat modules, so by Corollary 3.92 M is
flat.
(ii) =⇒ (i): see [Ei, Thm. A6.6]. 
11.1. Flat Base Change.
Proposition 3.102. (Stability of flatness under base change) Let M be a flat
R-module, and f : R → S a ring homomorphism. Then S ⊗R M is a flat S-module.
Exercise 3.89. Prove Proposition 3.102.
Exercise 3.90. Show: the tensor product of flat R-modules is a flat R-module.
86 3. MODULES

Exercise 3.91. Let R be a nonzero commutative ring, and n, m ∈ N.


a) Show that Rm ∼ = Rn iff m = n.
b) Suppose that ϕ : Rm → Rn is a surjective R-module map. Show that m ≥ n.
c)12 Suppose that ϕ : Rm → Rn is an injective R-module map. Show that m ≤ n.
d) Find a noncommutative ring R for which part a) fails.
Theorem 3.103. (Hom commutes with flat base change) Let S be a flat R-
algebra and M, N R-modules with M finitely presented. Then the canonical map
ΦM : S ⊗R HomR (M, N ) → HomS (M ⊗R S, N ⊗R S)
induced by (s, f ) 7→ (m ⊗ t) 7→ f (m) ⊗ st is an isomorphism.
Proof. (Hochster) It is immediate that ΦR is an isomorphism and that ΦM1 ⊕M2 =
ΦM1 ⊕ ΦM2 , and thus ΦM is an isomorphism when M is finitely generated free. For
finitely presented M , there is an exact sequence
H→G→M →0
with H and G finitely generated free modules. Now we have the following commu-
tative diagram:
0 −→ 0
M θ
S ⊗R HomR (M, N ) −→ HomS (M ⊗R S, N ⊗R S)
G θ
S ⊗R HomR (G, N ) −→ HomS (G ⊗R S, N ⊗R S)
H θ
S ⊗R HomR (H, N ) −→ HomS (H ⊗R S, N ⊗R S).
The right column is obtained by first applying the exact functor A 7→ A ⊗R S and
then applying the right exact cofunctor U 7→ HomS (U, N ⊗R S), so it is exact.
Similarly, the left column is obtained by first applying the right exact cofunctor
A 7→ HomR (A, N ) and then applying the exact (since R is flat) functor A 7→ A⊗R S,
so is exact. Since G and H are finitely generated free, θG and θH are isomorphisms,
and a diagram chase shows that θM is an isomorphism. 

12. Faithful flatness


Proposition 3.104. For an R-module M , the following are equivalent:
(i) For a sequence
α β
(11) N1 −→ N2 −→ N3
of left R-modules to be exact it is necessary and sufficient that
A B
(12) M ⊗R N1 −→ M ⊗R N2 −→ M ⊗R N3
be exact.
(ii) M is flat and for all nonzero R-modules N , M ⊗R N 6= 0.
(iii) M is flat and for all nonzero R-module maps u : N → N 0 ,
1M ⊗ u : M ⊗R N → M ⊗R N 0 is not zero .
(iv) M is flat and for every m ∈ MaxSpec R, mM ( M .
(v) M is flat and for every p ∈ Spec R, pM ( M .
A module satisfying these equivalent conditions is faithfully flat.
12This is actually quite challenging.
12. FAITHFUL FLATNESS 87

Proof. (i) =⇒ (ii): Certainly (i) implies that M is flat. Moreover, if N is


a nonzero R-module such that M ⊗ N = 0, then 0 → N → 0 is not exact but its
tensor product with M is exact, contradicting (i).
(ii) =⇒ (iii): Let I = Im(u); then M ⊗ I = Im(1M ⊗ u). So assuming (ii) and
that I 6= 0, we conclude Im(1M ⊗ u) 6= 0.
(iii) =⇒ (i): Assume (iii). Then, since M is flat, if (11) is exact, so is (12).
Conversely, suppose (12) is exact, and put I = Im(α), K = ker(β). Then B ◦ A =
1M ⊗ (β ◦ α) = 0, so β ◦ α = 0, or in other words, I ⊂ K. We may therefore form
the exact sequence
0 → I → K → K/I → 0,
and tensoring with the flat module M gives an exact sequence
0 → M ⊗ I → M ⊗ K → M ⊗ K/I → 0.
But M ⊗ K = M ⊗ I by hypothesis, so K/I = 0 and I = K.
(ii) =⇒ (iv): Let m ∈ MaxSpec R. Then R/m is a nonzero R-module, so by (ii)
so is M ⊗ R/m = M/mM , i.e., mM ( M .
(iv) =⇒ (ii): Assume (iv) holds. Then, since every proper ideal is contained in
a maximal ideal, we have moreover that for all proper ideals I of R, IM ( M , or
equivalently M ⊗ (R/I) 6= 0. But the modules of the form R/I as I ranges over
all proper ideals of R are precisely all the cyclic (a.k.a. monogenic) R-modules, up
to isomorphism. Now if N is any nonzero R-module, choose 0 6= x ∈ M and let
N 0 = hxi by the cyclic submodule spanned by x. It follows that M ⊗ N 0 6= 0. Since
M is flat, N 0 ,→ N implies M ⊗ N 0 ,→ M ⊗ N , so M ⊗ N 6= 0.
(iv) ⇐⇒ (v): this follows immediately from the proofs of the last two implications,
as we leave it to the reader to check. 
Exercise 3.92. Show: (iv) ⇐⇒ (v) in Proposition 3.104.
Corollary 3.105. Let M be a faithfully flat and u : N → N 0 an R-module
map. Then:
a) u is injective iff 1M ⊗ u : M ⊗ N → M 0 ⊗ N is injective.
b) u is surjective iff 1M ⊗ u is surjective.
c) u is an isomorphism iff 1M ⊗ u is an isomorphism.
Exercise 3.93. Deduce Corollary 3.105 from Proposition 3.104.
Exercise 3.94. Use each of the criteria of Proposition 3.104 to show that the
(flat) Z-module Q is not faithfully flat.
Exercise 3.95. Show: a faithfully flat module is faithful and flat, and that –
unfortunately! – a flat, faithful module need not be faithfully flat.
Exercise 3.96. Show: a nonzero free module is faithfully flat but that a nonzero
(even finitely generated) projective module need not be.

L Exercise 3.97. Let {Mi }i∈I be a family of flat R-modules, and put M =
i∈I Mi .
a) Suppose that for some i, Mi is faithfully flat. Show that M is faithfully flat.
b) Give an example where no Mi is faithfully flat yet M is faithfully flat.
Proposition 3.106. Let 0 → M 0 → M → M 00 → 0 be a short exact sequence
of R-modules. Suppose M 0 and M 00 are flat and that at least one is faithfully flat.
Then M is faithfully flat.
88 3. MODULES

Proof. By Proposition 3.94, M is flat. Now let N be any R-module. Since


M 00 is flat, Tor1 (M 00 , N ) = 0 so
0 → M 0 ⊗ N → M ⊗ N → M 00 ⊗ N → 0
is exact. Thus if M ⊗ N = 0 then M 0 ⊗ N = M 00 ⊗ N = 0. Since one of M 0 , M 00 is
faithfully flat, by criterion (ii) of Proposition 3.104 we have N = 0, and then that
same criterion shows that M is faithfully flat. 

By a faithfully flat R-algebra, we mean a ring S equipped with a ring homo-


morphism R → S making S into a faithfully flat R-module.
Proposition 3.107. Let f : R → S be a ring map and M an R-module. Then:
M is faithfully flat iff M ⊗R S is faithfully flat.
Proof. The key fact is that for any S-module N , we have
(M ⊗R S) ⊗S N ∼
=R M ⊗R N.
With this, the proof becomes straightforward and is left to the reader. 

Exercise 3.98. Complete the proof of Proposition 3.107.


Theorem 3.108. For a flat algebra f : R → S, the following are equivalent:
(i) S is faithfully flat over R.
(ii) f ∗ : MaxSpec S → MaxSpec R is surjective.
(iii) f ∗ : Spec S → Spec R is surjective.
Proof. (i) ⇐⇒ (ii): Let m be any maximal ideal of R. Then mS ( S holds
iff there is a maximal ideal M of S containing mS iff f ∗ (M) = m. The equivalence
now follows from criterion (iv) of Proposition 3.104.
(i) =⇒ (iii): Let p ∈ Spec R, and let k(p) be the fraction field of the domain
R/p. By faithful flatness, S ⊗R k(p) is a nonzero k(p)-algebra so has a prime
f g
ideal P. Consider the composite map h : R → S → S ⊗R k(p). We claim that
g ∗ : Spec(S ⊗R k(p)) → Spec R has image precisely {p}. The proof of this result, a
spectral description of the fiber of the morphism f : R → S over p, will have to
wait until we have developed the theory of localization in §7.3. Assuming it for now,
we get that g ∗ (P) is a prime ideal of Spec S such that f ∗ g ∗ (P) = (g ◦ f )∗ (P) = p,
so f ∗ : Spec S → Spec R is surjective.
(iii) =⇒ (ii): Let m ∈ MaxSpec R ⊂ Spec R. By assumption, the set of prime
ideals P of S such that f ∗ P = m is nonempty. Moreover the union of any chain of
prime ideals pulling back to m is again a prime ideal pulling back to p, so by Zorn’s
Lemma there exists an ideal M which is maximal with respect to the property that
f ∗ M = m. Suppose M is not maximal and let M0 be a maximal ideal properly
containing M. Then by construction f ∗ (M0 ) properly contains the maximal ideal
m of R, i.e., f ∗ (M)0 ) = R, contradicting the fact that prime ideals pull back to
prime ideals. So M is indeed maximal in S. 

Proposition 3.109. Let f : R ,→ S be a ring extension such that S is a


faithfully flat R-module, and let M be an R-module. Then:
a) M is finitely generated iff M ⊗R S is finitely generated.
b) M is finitely presented iff M ⊗R S is finitely presented.
12. FAITHFUL FLATNESS 89

Proof. Note first that the properties of finite generation and finite presen-
tation are preserved by arbitrary base change f : R → S. So it suffices to prove
that if M ⊗R S is finitely generated (resp. finitely presented), then M is finitely
generated (resp. finitely presented).
a) Since M ⊗R S is finitely generated over S, it has a finite set of S-module gen-
erators of the form xi ⊗ 1. Let N = hx1 , . . . , xn iR and ι : N ,→ M the canonical
injection. Then ιS : N ⊗R S → M ⊗R S is an isomorphism, so by faithful flatness
ι was itself an isomorphism and thus M = hx1 , . . . , xn i is finitely generated.
b) By part a), M is finitely generated over R, so let u : Rn → M be a surjection.
Since M ⊗R S is finitely presented, the kernel of uS : S n → M ⊗R S is finitely
generated over S. Since by flatness ker uS = (ker u)S , part a) shows that ker u is
finitely generated and thus that M is finitely presented. 
Lemma 3.110. Let f : R → S be a ring map, and let M, N be R-modules.
a) There is a canonical S-module map
ω : HomR (M, N ) ⊗R S → HomS (M ⊗R S, N ⊗R S)
such that for all u ∈ HomR (M, N ), ω(u ⊗ 1) = u ⊗ 1B .
b) If S is flat over R and M is finitely generated, then ω is injective.
c) If S is flat over R and M is finitely presented, then ω is an isomorphism.
Exercise 3.99. Prove Lemma 3.110. (It is not difficult, really, but it is some-
what technical. Feel free to consult [B, p. 23] for the details.)
Theorem 3.111. (Faithfully flat descent for projective modules) Let f : R ,→ S
be a faithfully flat ring extension, and let P be an R-module. Then P is finitely
generated and projective iff P ⊗R S is finitely generated and projective.
Proof. Begin, once again the implication P finitely generated projective im-
plies P ⊗R S is finitely generated projective holds for any base change. So suppose
P ⊗R S is finitely generated projective. Then P ⊗R S is finitely presented, so by
Proposition 3.109, M is finitely presented. It remains to show that M is projective.
Let v : M → N be a surjection of R-modules. We wish to show that the natu-
ral map HomR (P, M ) → HomR (P, N ) is surjective. Because of the faithful flatness
of S/R, it is sufficient to show that HomR (P, M ) ⊗R S → HomR (P, N ) ⊗R S is
surjective, and by Lemma 3.110 this holds iff
HomS (P ⊗R S, M ⊗R S) → HomS (P ⊗R S, N ⊗R S)
is surjective. But this latter map is surjective because M ⊗R S → N ⊗R S is
surjective and the S-module P ⊗R S is projective by assumption. 
CHAPTER 4

First Properties of Ideals in a Commutative Ring

1. Introducing maximal and prime ideals


Consider again the set I(R) of all ideals of R, partially ordered by inclusion. The
maximal element is the ideal R itself, and the minimal element is the ideal (0).

In general our attitude to the ideal R of R is as follows: although we must grudg-


ingly admit its existence – otherwise, given a subset S of R it would be in general
a difficult question to tell whether the ideal hSi generated by S “exists” (i.e., is
proper) or not – nevertheless we regard it as exceptional and try to ignore it as
much as possible. Because of this we define an ideal I of R to be maximal if it is
maximal among all proper ideals of R, i.e., I ( R and there does not exist J such
that I ( J ( R. That this is a more interesting concept than the literally maximal
ideal R of R is indicated by the following result.
Proposition 4.1. For an ideal I of R, the following are equivalent:
(i) I is maximal.
(ii) R/I is a field.
Proof. Indeed, R/I is a field iff it has precisely two ideals, I and R, which by
the Correspondence Theorem says precisely that there is no proper ideal strictly
containing I. 

Example: In R = Z, the maximal ideals are those of the form (p) for p a prime
number. The quotient Z/pZ is the finite field of order p.

Does every ring have a maximal ideal? With a single (trivial) exception, the answer
is yes, assuming – as we must, in order to develop the theory as it is used in other
branches of mathematics – suitable transfinite tools.
Proposition 4.2. Let R be a nonzero ring and I a proper ideal of R. Then
there exists a maximal ideal of R containing I.
Proof. Consider the set S of all proper ideals of R containing I, partially
ordered by inclusion. Since I ∈ S, S is nonempty. Moreover the union of a chain
of ideals is an ideal, and the union of a chain of proper ideals is proper (for if 1
were in the union, it would have to lie in one of the ideals of the chain). Therefore
by Zorn’s Lemma we are entitled to a maximal element of S, which is indeed a
maximal ideal of R that contains I. 

Corollary 4.3. A nonzero ring R contains at least one maximal ideal.


Proof. Apply Proposition 4.2 with I = (0). 
91
92 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Remark: The zero ring has the disquieting property of having no maximal ideals.

Remark: The appeal to Zorn’s Lemma cannot be avoided, in the sense that Corol-
lary 4.3 implies the Axiom of Choice (AC). In fact, W. Hodges has shown that the
axioms of ZF set theory together with the statement that every UFD (see §15) has
a maximal ideal already implies AC [Ho79].

A proper ideal I of a ring R is prime if xy ∈ I implies x ∈ I or y ∈ I.


Exercise 4.1. Let p be a prime ideal of R.
a) Suppose x1 , . . . , xn are elements of R such that x1 · · · xn ∈ p. Then xi ∈ p for
some at least one i.
b) In particular, for x ∈ R and n ∈ Z+ we have xn ∈ p, then x ∈ p.
Proposition 4.4. Let f : R → S be a homomorphism of rings, and let J be
an ideal of S.
a) Put f ∗ (J) := f −1 (J) = {x ∈ R | f (x) ∈ J}. Then f ∗ (J) is an ideal of R.
b) If J is a prime ideal, so is f ∗ (J).
Exercise 4.2. Prove Proposition 4.4.
Proposition 4.5. For a commutative ring R, the following are equivalent:
(i) If x, y ∈ R are such that xy = 0, then x = 0 or y = 0.
(ii) If 0 6= x ∈ R and y, z ∈ R are such that xy = xz, then y = z.
A ring satisfying either of these two properties is called a domain.
Proof. Assume (i), and consider xy = xz with x 6= 0. We have x(y − z) = 0,
and since x 6= 0, (i) implies y − z = 0, i.e., y = z. Assuming (ii) suppose xy = 0
with x 6= 0. Then xy = 0 = x · 0, so applying cancellation we get y = 0. 
A zero divisor in a ring R is an element x such that there exists 0 6= y ∈ R with
xy = 0. (In particular 0 is a zero divisor, albeit not a very interesting one.) So
property (i) expresses that there are no zero divisors other than 0 itself. Property
(ii) makes sense in any commutative monoid and is called cancellation.
Proposition 4.6. For an ideal I in a ring R, the following are equivalent:
(i) I is prime.
(ii) R/I is a domain.
Exercise 4.3. Prove Proposition 4.6.
Corollary 4.7. A maximal ideal is prime.
Proof. If I is maximal, R/I is a field, hence a domain, so I is prime. 
Corollary 4.7 is the first instance of a somewhat mysterious meta-principle in ideal
theory: for some property P of ideals in a ring, it is very often the case that an
ideal which is maximal with respect to the satisfaction of property P (i.e., is not
strictly contained in any other ideal satisfying P) must be prime. In the above, we
saw this with P = “proper”. Here is another instance:
Proposition 4.8. (Multiplicative Avoidance) Let R be a ring and S ⊂ R.
Suppose: 1 is in S; 0 is not in S; and S is closed under multiplication: S · S ⊂ S.
Let IS be the set of ideals of R which are disjoint from S. Then:
a) IS is nonempty;.
2. RADICALS 93

b) Every element of IS is contained in a maximal element of IS .


c) Every maximal element of IS is prime.
Proof. a) (0) ∈ IS . b) Let I ∈ IS . Consider the subset PI of IS consisting
of ideals which contain I. Since I ∈ PI , PI is nonempty; moreover, any chain in PI
has an upper bound, namely the union of all of its elements. Therefore by Zorn’s
Lemma, PI has a maximal element, which is clearly also a maximal element of IS .
c) Let I be a maximal element of IS ; suppose that x, y ∈ R are such that xy ∈ I.
If x is not in I, then hI, xi ) I and therefore contains an element s1 of S, say
s1 = i1 + ax.
Similarly, if y is not in I, then we get an element s2 of S of the form
s2 = i2 + by.
But then
s1 s2 = i1 i2 + (by)i1 + (ax)i2 + (ab)xy ∈ I ∩ S,
a contradiction. 

In fact Corollary 4.7 is precisely the special case S = {1} of Proposition 4.8.

If I and J are ideals of R, we define the product IJ to be the ideal generated


by allPelements of the form xy with x ∈ I, y ∈ J. Every element of IJ is of the
n
form i=1 xi yi with x1 , . . . , xn ∈ I, y1 , . . . , yn ∈ J.

The following simple result will be used many times in the sequel.
Proposition 4.9. Let p be a prime ideal and I1 , . . . , In be ideals of a ring R.
If p ⊃ I1 · · · In , then p ⊃ Ii for at least one i.
Proof. An easy induction argument reduces us to the case of n = 2. So
suppose for a contradiction that p ⊃ I1 I2 but there exists x ∈ I1 \ p and y ∈ I2 \ p.
Then xy ∈ I1 I2 ⊂ p; since p is prime we must have x ∈ p or y ∈ p, contradiction. 

Exercise 4.4. Show: Proposition 4.9 characterizes prime ideals, in the sense
that if p is any ideal such that for all ideals I, J of R, p ⊂ IJ implies p ⊂ I or
p ⊂ JJ, then p is a prime ideal.
For an ideal I and n ∈ Z+ , we denote the n-fold product of I with itself by I n .
Corollary 4.10. If p, I are ideals and p is prime, then p ⊃ I n =⇒ p ⊃ I.

2. Radicals
An element x of a ring R is nilpotent if xn = 0 for some n ∈ Z+ . Obviously
0 is a nilpotent element; a ring in which 0 is the only nilpotent element is called
reduced. An ideal I of R is nil if every element of I is nilpotent. An ideal I is
nilpotent if there exists n ∈ Z+ such that I n = (0).
Proposition 4.11. Let I be an ideal of a ring R.
a) If I is nilpotent, then I is a nil ideal.
b) If I is finitely generated and nil, then I is nilpotent.
94 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Proof. Part a) is immediate from the definition.


Suppose I = ha1 , . . . , an iR . Since I is nil, for each i, 1 ≤ i ≤ r, there exists ni
such that ani i = 0. Let N = n1 + . . . + nr . We claim I N = 0. Indeed, an arbitrary
element of I is of the form x1 a1 + . . . + xn an . Raising this element to Pthe nth power
r
yields a sum of monomials of the form xj11 · · · xjrr aj11 · · · ajrr , where i=1 ji = N . If
we had for all i that ji < ni , then certainly j1 + . . . + jr < N . So for at least one i
we have ji ≥ ni and thus xji i = 0; so every monomial term equals zero. 
Exercise 4.5. The bound on the nilpotency index of ha1 , . . . , an i we gave in
the preceding argument is a bit lazy: we can do better.
a) Let a1 , . . . , ar be elements of a ring and let n1 , . . . , nr ∈ Z+ be such that
an1 1 = . . . = anr r = 0. Show:
ha1 , . . . , ar in1 +...+nr −(r−1) = (0).
b) Show that the bound of part a) is best possible for all positive integers
r, n1 , . . . , nr .
(Suggestion: for a field k, consider k[t1 , . . . , tr ]/htn1 1 , . . . , tnr r i.)
Exercise 4.6. Find a ring R and an ideal I of R which is nil but not nilpotent.
The nilradical N of R is the set of all nilpotent elements of R.
Proposition 4.12. Let R be a ring.
a) The nilradical N is a nil ideal of R.
b) The quotient R/N is reduced.
c) The map q : R → R/N is universal for maps from R into a reduced ring.
d) The nilradical is the intersection of all prime ideals of R.
Proof. a) It suffices to show that N is an ideal. The only part of this that
is not absolutely immediate is closure under addition. But by Proposition 4.11 the
ideal generated by two nilpotent elements is nilpotent, hence nil, hence consists of
nilpotent elements, so in particular the sum of two nilpotent elements is nilpotent.
b) Let r + N be a nilpotent element of R/N , so there exists n ∈ Z+ such that
rn ∈ N . But this means there exists m ∈ Z+ such that 0 = (rn )m = rnm , and thus
r itself is a nilpotent element.
c) In plainer terms: if S is a reduced ring and f : R → S is a ring homomorphism,
then there exists a unique homomorphism f : R/N → S such that f = f ◦ q. Given
this, the proof is straightforward, and we leave it to the reader.
d) Suppose x is a nilpotent element of R, i.e., ∃n ∈ Z+ such that xn = 0. If T p is
a prime ideal, then since 0 = x · · · x ∈ p, we conclude x ∈ p: this shows N ⊂ p.
Conversely, suppose x is not nilpotent. Then the set Sx := {xn | n ∈ N} satisfies
(i) and (ii) of Proposition 5.24, so we may apply that result to get a prime ideal p
which is disjoint from Sx , hence not containing x. 
Exercise 4.7. Prove Proposition 4.12c).
Exercise 4.8. Let f : R → T be a ring homomorphism. Show:
f∗ (nil R) ⊂ nil T.
For a ring R, what are the units in the polynomial ring R[t]? If R is a domain, then
the leading coefficient of any nonzero polynomial is nonzero hence not a zero-divisor.
This implies that for nonzero f, g ∈ R[t] we have deg(f g) = deg(f ) + deg(g), from
2. RADICALS 95

which it follows that every unit has degree zero, i.e., is a constant polynomial and
then that R[t]× = R× . The following result treats the general case by reduction to
this case.
Proposition 4.13. Let R be a nonzero ring, and let R[t] be the polynomial
ring over R. Let f be a nonzero element of R[t] and write
f = an tn + . . . + a1 t + a0 , an 6= 0.
Then f ∈ R[t]× iff a0 ∈ R× and for all 1 ≤ i ≤ n we have ai ∈ nil R.
Proof. Put g := an tn + . . . + a1 t.
First suppose that a0 ∈ R× and ai ∈ nil R for all 1 ≤ i ≤ n. Then is g nilpotent
by Exercise 4.8, so f = g + a0 is the sum of a nilpotent element and a unit. If m is
any maximal ideal of R[t] then if f ∈ m then a0 = f − g ∈ m, contradiction. So f
lies in no maximal ideal, hence is a unit of R[t].
Now suppose that f ∈ R[t]× . For a prime ideal p of R, let
qp : R[t] → R[t]/(pR[t]) ∼= (R/p)[t]
be the quotient map. Then qp (f ) is a unit in (R/p)[t], which implies that ai ∈ p for
all 1 ≤ i ≤ n. Since this holds for all prime ideals p we have ai ∈ nil R. Moreover
qp (a0 ) ∈ (R/p)× , hence a0 ∈
/ p. As above, this means that a0 lies in no maximal
ideal, hence a0 ∈ R× . 
Proposition 4.14. (Lifting Idempotents Modulo a Nil Ideal) Let R be a ring,
let N be a nil ideal of R, and let e = e2 be an idempotent element of R/N . Then
there is a unique idempotent e of R such that e (mod N ) = e.
Proof. (Jacobson [J2, Prop. 7.14])
Step 1: We will prove the existence of e.1 First let x ∈ R be such that x
(mod N ) = e. Then z = x − x2 is nilpotent: there is n ∈ Z+ such that z n = 0. Put
y = 1 − x. Then
0 = z n = (x(1 − x))n = xn y n ,
so
1 = 12n−1 = (x + y)2n−1 = e + f,
where e is a sum of terms xi y 2n−1−i with n ≤ i ≤ 2n − 1 and f is a sum of the
terms xi y 2n−1−i with 0 ≤ i ≤ n − 1. Since xn y n = 0, any term in e annihilates any
term in f . Hence ef = 0 = f e. Since e + f = 1, we have
e = e(e + f ) = e2 + ef = e2 , f = (e + f )f = ef + f 2 = f 2 .
Every term in e except x2n−1 contains the factor xy = z, so x2n−1 ≡ e (mod N ).
Since x ≡ x2 ≡ x3 ≡ . . . ≡ x2n−1 (mod N ), we have e ≡ x ≡ e (mod N ).
Step 2: Let e, z ∈ R with e idempotent, z nilpotent and e + z idempotent. Then
e + z = (e + z)2 = e2 + 2ez + z 2 = e + 2ez + z 2
so
z 2 = (1 − 2e)z.
It follows that
z 3 = (1 − 2e)z 2 = (1 − 2e)2 z = (1 − 4e + 4e2 )z = (1 − 4e + 4e)z = z

1This step holds for not-necessarily-commutative rings.


96 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

and thus
z 2k+1 = z ∀k ∈ Z+ .
Since z is nilpotent, it follows that z = 0. 
An ideal I of a ring R is radical if for all x ∈ R, n ∈ Z+ , xn ∈ I implies x ∈ I.
Exercise 4.9. a) Show: a prime ideal is radical.
b) Exhibit a radical ideal which is not prime.
c) Find all radical ideals in R = Z.
d) Show: R is reduced iff (0) is a radical ideal. T
e) Let {Ii } be a set of radical ideals in a ring R. Show I = i Ii is a radical ideal.
Exercise 4.10. Let p1 and p2 be prime ideals of a ring R. By Exercise 4.9,
we have p1 ∩ p2 is a radical ideal.
a) Show: if p1 + p2 = R then p1 p2 is radical.
b) Give an example in which p1 6= p2 and p1 p2 is not radical.
For any ideal I of R, we define the radical of I:
r(I) = {x ∈ R | ∃n ∈ Z+ xn ∈ I}.
Proposition 4.15. Let R be a commutative ring and I, J ideals of R.
a) r(I) is the intersection of all prime ideals containing I, and is a radical ideal.
b) (i) I ⊂ r(I), (ii) r(r(I)) = r(I); (iii) I ⊂ J =⇒ r(I) ⊂ r(J).
c) r(IJ) = r(I ∩ J) = r(I) ∩ r(J).
d) r(I + J) = r(r(I) + r(J)).
e) r(I) = R ⇐⇒ I = R.
f ) For all n ∈ Z+ , r(I n ) = r(I).
g) If J is finitely generated and r(I) ⊃ J, then there is n ∈ Z+ such that I ⊃ J n .
Proof. Under the canonical homomorphism q : R → R/I, r(I) = q −1 (N (R/I)).
By Proposition 4.4a), r(I) is an ideal.
a) Since N is the intersection of all prime ideals of R/I, r(I) is the intersection of
all prime ideals containing I, which is, by Exericse 4.9e), a radical ideal.
b) (i) is immediate from the definition, and (ii) and (iii) follow from the character-
ization of r(I) as the intersection of all radical ideals containing I.
c) Since IJ ⊂ I ∩ J, r(IJ) ⊂ r(I ∩ J). If xn ∈ I ∩ J, then x2n = xn xn ∈ IJ, so
x ∈ r(IJ); therefore r(IJ) = r(I ∩ J). Since I ∩ J is a subset of both I and J,
r(I ∩ J) ⊂ r(I) ∩ r(J). Conversely, if x ∈ r(I) ∩ r(J), then there exist m and n
such that xm ∈ I and xn ∈ J, so xmn ∈ I ∩ J and x ∈ r(I ∩ J).
d) Since I + J ⊂ r(I) + r(J), r(I + J) ⊂ r(r(I) + r(J)). A general element of
r(I)+r(J) is of the form x+y, where xm ∈ I and y n ∈ J. Then (x+y)m+n ∈ I +J,
so x + y ∈ r(I + J).
e) Evidently r(R) = R. Conversely, if r(I) = R, then there exists n ∈ Z+ such that
1 = 1n ∈ I.
f) By part a), r(I n ) is the intersection of all prime ideals p ⊃ I n . But by Corollary
4.10, a prime contains I n iff it contains I, so r(I n ) = r(I).
g) Replacing R with R/I we may assume I = 0. Then J is a finitely generated nil
ideal, so it is nilpotent. 
Remark: Proposition 4.15b) asserts that the mapping I 7→ r(I) is a closure op-
erator on the lattice I(R) of ideals of R.
2. RADICALS 97

An ideal p of a ring R is primary if every zero divisor of R/p is nilpotent. Equiva-


/ p =⇒ y n ∈ p for some n ∈ Z+ . More on primary ideals in §10.3.
lently, xy ∈ p, x ∈

We also define the Jacobson radical J(R) as the intersection of all maximal
ideals of R. Evidently we have N ⊂ J(R).
Proposition 4.16. Let R be a ring. An element x of R lies in the Jacobson
radical J(R) iff 1 − xy ∈ R× for all y ∈ R.
Proof. Suppose x lies in every maximal ideal of R. If there is y such that
1 − xy ∈/ R× , then 1 − xy lies in some maximal ideal m, and then x ∈ m implies
xy ∈ m and then 1 = (1 − xy) + xy ∈ m, a contradiction. Conversely, suppose that
there is a maximal ideal m of R which does not contain x. Then hm, xi = R, so
1 = m + xy for some m ∈ m and y ∈ R, and thus 1 − xy is not a unit. 
Proposition 4.17. Let J be an ideal of R contained in the Jacobson radical,
and let ϕ : R → R/J be the natural map.
a) For all x ∈ R, x ∈ R× ⇐⇒ ϕ(x) ∈ (R/J)× : ϕ is unit-faithful.
b) The map ϕ× : R× → (R/J)× is surjective.
Proof. a) For any homomorphism of rings ϕ : R → S, if x ∈ R× then there is
y ∈ R with xy = 1, so 1 = ϕ(1) = ϕ(xy) = ϕ(x)ϕ(y), and thus ϕ(x) ∈ S × . For the
converse we assume S = R/J and let x ∈ R be such that ϕ(x) ∈ (R/J)× . Then
there is y ∈ R such that xy−1 ∈ J. Thus for each maximal ideal m of R, xy−1 ∈ m.
It follows that x ∈
/ m, for otherwise xy ∈ m and thus 1 = xy − (xy − 1) ∈ m. So x
is not contained in any maximal ideal and thus x ∈ R× .
b) This is immediate from part a): in fact we’ve shown that every preimage under
ϕ of a unit in R/J is a unit in R. 
Exercise 4.11. For a ring R, let ι : R ,→ R[t] be the inclusion into the
polynomial ring R[t].
a) Show:
J(R[t]) = ι∗ (nil R).
b) Thus J(R[t]) = (0) if R is a domain. Give a simpler direct proof of this.
Remark: It is not yet clear why we have defined these two different notions of
“radical.” Neither is it so easy to explain in advance, but nevertheless let us make
a few remarks. First, the Jacobson radical plays a very important role in the theory
of noncommutative rings, especially that of finite dimensional algebras over a field.
(Indeed, a finite dimensional k-algebra is semisimple – i.e., a direct product of
algebras without nontrivial two-sided ideals – iff its Jacobson radical is zero. In the
special case of commutative algebras this comes down to the simpler result that
a finite dimensional commutative k-algebra is reduced iff it is a product of fields.)
One important place in commutative algebra in which the Jacobson radical J(R)
appears – albeit not by name, because of the necessity of putting the results in
a fixed linear order – is in the statement of Nakayama’s Lemma. In general, the
defining condition of nil(R) – i.e., as the intersection of all prime ideals of R –
together with the fact that the radical of an arbitrary ideal I corresponds to the
nilradical of R/I, makes the nilradical more widely useful in commutative algebra
(or so it seems to the author of these notes). It is also important to consider
98 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

when the nil and Jacobson radicals of a ring coincide. A ring R for which every
homomorphic image S has nil(S) = J(S) is called a Jacobson ring; such rings
will be studied in detail in §12.

3. Comaximal ideals
Two ideals I and J in a ring R are comaximal if I + J = R. A family of ideals in
R is pairwise comaximal if any two members of the family are comaximal.
Pn Q
Exercise 4.12. Let I1 , . . . , In be pairwise comaximal. Show: j=1 i6=j Ii =
R.
Proposition 4.18. Let I and J be ideals in R. If r(I) and r(J) are comaximal,
so are I and J.
Proof. Apply Proposition 4.15d) and 4.15e) to r(I) + r(J) = R:
R = r(r(I) + r(J)) = r(I + J) = I + J. 
An immediate corollary of Proposition 4.18 is that if {Ii } are pairwise comaximal
and {ni } are any positive integers, then {Iini } are pairwise comaximal.
Lemma 4.19.
Tn Let K1 , . . . , Kn be pairwise comaximal ideals in the ring R. Then
K1 · · · Kn = i=1 Ki .
Proof. We go by induction on n: n = 1 is trivial and n = 2 is Lemma 3.17b).
Tnthe theorem is true for0 any family of n − 1 pairwise comaximal ideals.
Suppose Let
K 0 = i=2 Ki ; by T
induction, K = K2 · · · Kn . By Lemma 3.17c), K1 + K 0 = R, so
n
by the n = 2 case i=1 Ki = K1 ∩ K 0 = K1 K 0 = K1 · · · Kn . 
Theorem 4.20. (Chinese Remainder Theorem, or “CRT”) Let R be a ring
and I1 , . . . , In a finite set of pairwise comaximal ideals. Consider the natural map
n
Y
Φ:R→ R/Ii ,
i=1
x 7→ (x + Ii )ni=1 .
Then Φ is surjective with kernel I1 · · · In , so that there is an
induced isomorphism
n

Y
(13) Φ : R/(I1 · · · In ) → R/Ii .
i=1
Tn
Proof. The map Φ is well-defined T and has kernel i=1 Ii . Since the Ii ’s are
n
pairwise comaximal, Lemma 4.19 gives i=1 Ii = I1 · · · In . So it remains to show
that Φ is surjective. We prove this by induction on n, the Q case n = 1 being trivial.
0 0 n−1
So we may assume that the natural map Φ : R → R := i=1 R/Ii is surjective,
with kernel I 0 := I1 · · · In−1 . Let (r0 , s) be any element of R0 ×R/In . By assumption,
there exists r ∈ R such that Φ0 (r + I 0 ) = r0 . Let s be any element of R mapping
to s ∈ R/In . By Lemma 3.17c) we have I 0 + In = R, so there exist x ∈ I 0 , y ∈ In
such that s − r = x + y. Then Φ0 (r + x) = r0 , and r + x ≡ r + x + y ≡ s (mod In ),
so Φ(r + x) = (r0 , s) and Φ is surjective. 
Remark: In the classical case R = Z, we can write Ii = (ni ) and then we are trying
to prove – under the assumption that the ni ’s are coprime in pairs in the sense of
elementary number theory – that the injective ring homomorphism Z/(n1 · · · nn ) →
Z/n1 × · · · × Z/nn is an isomorphism. But both sides are finite rings of order
3. COMAXIMAL IDEALS 99

n1 · · · nn , so since the map is an injection it must be an isomorphism! Nevertheless


the usual proof of CRT in elementary number theory is much closer to the one we
gave in the general case: in particular, it is constructive.

QExercise 4.13 (Converse to CRT). Let I1 , . . . , In be ideals in a ring R. Show:


n
if i=1 R/Ii is a cyclic R-module, then I1 , . . . , In are pairwise comaximal.
The following modulization of CRT is sometimes useful.
Theorem 4.21. (Module-theoretic CRT) Let R be a ring, I1 , . . . , In a finite
set
Tr of pairwise comaximal ideals, and let M be an R-module. Then (I1 · · · In )M =
i=1 i M , and there is an induced R-module isomorphism
I
n
Y
(14) ΦM : M/(I1 · · · Ir )M → M/Ii M.
i=1

Proof. Indeed ΦM = Φ ⊗R M , so it is an isomorphism. Thus


r n
!
\ Y
Ii M = ker M → M/Ii M = (I1 · · · Ir )M. 
i=1 i=1

Exercise 4.14. Let R Q be a ring and I1 , . . . , In any finite sequence of ideals.


n
Consider the map Φ : R → i=1 R/Ii as in CRT.
Tn if the {Ii } are pairwise comaximal.
a) Show that Φ is surjective only
b) Show that Φ is injective iff i=1 Ii = (0).
Exercise 4.15. P a) Let G be a finite commutative group with exactly one element
z of order 2. Show: x∈G x = z.
b) Let G be a finite
P commutative group which does not have exactly one element of
order 2. Show: x∈G x = 0.
c) Prove the following result of Gauss (a generalization of Wilson’s Theorem):
let N ∈ Z+ , and put
Y
P (N ) = x.
x∈(Z/N Z)×

Then: P (N ) = ±1, and the minus sign holds iff N = 4 or is of the form pm or
2pm for an odd prime p and m ∈ Z+ .
d) For a generalization to the case of (ZK /A)× , where A is an ideal in the ring ZK
of integers of a number field K, see [Da09]. Can you extend Dalawat’s results to
the function field case?
Exercise 4.16. Let K be a field, and put R = K[t].
a) Let n1 , . . . , nk be a sequence of non-negative integers and {x1 , . . . , xk } a k-
element subset of K. For 1 ≤ i ≤ k, let ci0 , . . . , cini be a finite sequence of ni + 1
elements of k (not necessarily distinct). By applying the Chinese Remainder The-
orem, show that there is a polynomial P (t) such that for 1 ≤ i ≤ k and 0 ≤ j ≤ ni
we have P (j) (xi ) = cij , where P (j) (xi ) denotes the jth “formal” derivative of P
evaluated at xi . Indeed, find all such polynomials; what can be said about the least
degree of such a polynomial?
b) Use the proof of the Chinese Remainder Theorem to give an explicit formula for
such a polynomial P .
100 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Exercise 4.17. Let (M, ·) be a monoid and k be a field. A character on M


with values in k is a homomorphism of monoids from M to the multiplicative group
k × of k. Each character lies in the k-vector space k M of all functions from M to
k.
a) (Dedekind) Show: any finite set of characters is k-linearly independent.
b) Give an example of an infinite set of characters which is k-linearly dependent.
c) But what does this have to do with CRT? Well, the wikipedia article on CRT2
contains a proof of part a) using CRT. This is the proof I had in mind when I
originally wrote this exercise. But it seems to me now that this argument requires
M to be finite. Discuss.
Exercise 4.18. Show: for a ring R, the following are equivalent:
(i) R has finitely many maximal ideals.3
(ii) The quotient of R by its Jacobson radical J(R) is a finite product of fields.
We now give a commutative algebraic version of Euclid’s proof of the infinitude of
prime numbers. A special case for domains appears in [K, § 1.1, Exc. 8]. The case
in which R is infinite and R× is finite has appeared on an algebra qualifying exam
at UGA; the appearance of this unusually interesting and challenging problem on a
qual was remarked to me by both D. Lorenzini and B. Cook. I learned the stronger
version presented here from W.G. Dubuque.
Theorem 4.22. If R is infinite and #R > #R× , then MaxSpec R is infinite.
Proof. Since R is not the zero ring, it has at least one maximal ideal m1 .
We proceed by induction: given maximal ideals m1 , . . . , mn , we construct another
maximal ideal.
Step 1: Suppose J + 1 ⊂ R× . Then
#J = #(J + 1) ≤ #R× < #R.
Moreover, by Proposition 4.16, J is contained in the Jacobson radical of R and
thus by Proposition 4.17, R× → (R/J)× is surjective. It follows that #(R/J)× ≤
× ∼ Qn
#R < #R: by the Chinese remainder Theorem, R/J = i=1 R/mi , hence there
is an injection (R/mi )× → (R/J)× . Putting the last two sentences together we
conclude #(R/mi )× < #R, and thus, since R/mi is a field and R is infinite,
#R/mi = #R/mi + 1 < #R. Finally this gives
Yn
#R = #J · #R/J = #J · #R/mi < (#R)n+1 = #R,
i=1
a contradiction.
Step 2: So J + 1 6⊂ R× . Let x ∈ J + 1 \ R× , and let m be a maximal ideal containing
x. Then for all 1 ≤ i ≤ n, x − 1 ∈ J ⊂ mi , so 1 = x + (1 − x) ∈ m + mi . It follows
that m 6⊆ mi , so it’s a new maximal ideal.

In §15.11, following a discussion of factorization in domains, we will give a variant
on this result that gives a sufficient condition for a domain R to have infinitely
many pairwise nonassociate irreducible elements.
2https://en.wikipedia.org/wiki/Chinese_remainder_theorem#Dedekind.27s_theorem
3Such rings are typically called semilocal. I am not a fan of the terminology – it seems to
either suggest that R has one half a maximal ideal (whatever that could mean) or two maximal
ideals. But it is well entrenched, and I will not campaign to change it.
5. THE PRIME IDEAL PRINCIPLE OF LAM AND REYES 101

4. Local rings
Proposition 4.23. For a ring R, the following are equivalent:
(i) There is exactly one maximal ideal m.
(ii) The set R \ R× of nonunits forms a subgroup of (R, +).
(iii) The set R \ R× is a maximal ideal.
A ring satisfying these equivalent conditions is called a local ring.
Proof. Since R× = R \ m m, the union extending over all maximal ideals of
S
R, it follows that if there is only one maximal ideal m then m = R \ R× . This shows
(i) =⇒ (iii) and certainly (iii) =⇒ (ii). Conversely, since the set of nonunits of
a ring is a union of ideals, it is closed under multiplication by all elements of the
ring. Thus it is itself an ideal iff it is an additive subgroup: (ii) =⇒ (iii). The
implication (iii) implies (i) is very similar and left to the reader. 
Warning: In many older texts, a ring with a unique maximal ideal is called “quasi-
local” and a local ring is a Noetherian quasi-local ring. This is not our convention.

Local rings (especially Noetherian local rings) play a vital role in commutative al-
gebra: the property of having a single maximal ideal simplifies many ideal-theoretic
considerations, and many ring theoretic considerations can be reduced to the study
of local rings (via a process called, logically enough, localization: see Chapter 7).

A field is certainly a local ring. The following simple result builds on this triv-
ial observation to give some further examples of local rings:
Proposition 4.24. Let I be an ideal in the ring R.
a) If rad(I) is maximal, then R/I is a local ring.
b) In particular, if m is a maximal ideal and n ∈ Z+ then R/mn is a local ring.
T
Proof. a) We know that rad(I) = p⊃I p, so if rad(I) = m is maximal it must
be the only prime ideal containing I. Therefore, by correspondence R/I is a local
ring. (In fact it is a ring with a unique prime ideal.)
b) By Proposition 4.15f), r(mn ) = r(m) = m, so part a) applies. 
So, for instance, for any prime number p, Z/(pk ) is a local ring, whose maximal ideal
is generated by p. It is easy to see (using, e.g. the Chinese Remainder Theorem)
that conversely, if Z/(n) is a local ring then n is a prime power.
Example 4.25. The ring Zp of p-adic integers is a local ring. For any field
k, the ring k[[t]] of formal power series with coefficients in k is a local ring. Both
of these rings are also PIDs. A ring which is a local PID is called a discrete
valuation ring; these especially simple and important rings will be studied in
detail later.
Exercise 4.19. Show: a local ring is connected, i.e., e2 = e =⇒ e ∈ {0, 1}.

5. The Prime Ideal Principle of Lam and Reyes


A recurrent meta-principle in commutative algebra is that if F is a naturally given
family of ideals in commutative ring R, then it is often the case that every maximal
element of F is prime. In this section we review some known examples, give some
further classical ones, and then discuss a beautiful theorem of T.-Y. Lam and M.
102 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Reyes which gives a general criterion for this phemenon to occur.

Recall that for a ring R, I(R) is the monoid of ideals of R under multiplication.
For any F ⊂ I(R), let Max F denote the maximal elements of F (to be sure, this
means the elements of F which are not properly contained in any other element of
F, not the elements of F which are not contained in any other proper ideal!). We
say that F is an MP family if Max F ⊂ Spec R.
Example 4.26. a) Every maximal ideal is prime (Corollary 4.7), so the
family of all proper ideals in a ring R is an MP family.
b) If S ⊂ R is a multiplicative set, an ideal that is maximal with the property
of being disjoint from S is prime (Proposition 5.24), so the family of all
ideals in R that are disjoint from S is an MP family.
c) As we will see shortly, the family of all non-principal ideals in a ring R
is an MP family. Thus if every prime ideal is principal, every ideal is
principal: Theorem 4.29.
d) As we will see shortly, the family of all infinitely generated ideals in a ring
R is an MP family. This implies a result of Cohen (Theorem 4.30): if
every prime ideal is finitely generated, then R is Noetherian.
The challenge is to come up with a common explanation and proof for all of these
examples. One first observation is that there is a complementation phenomenon in
play here: for F ⊂ I(R), put F 0 = I(R) \ F. Then in each of the last three cases it
is most natural to view the MP family as F 0 for a suitable F: in the second case, F
is the set of ideals meeting S; in the third case, F is the set of all principal ideals;
in the fourth case F is the set of all finitely generated ideals.

Let us also recall that for I, J ∈ I(R),


(I : J) = {x ∈ R | xJ ⊂ I}.
For a, b ∈ R, we write (I : b) for (I : Rb) and (a : J) for (aR : J).
Exercise 4.20. For ideals I, J in R, show that
(15) (I : J)hI, Ji ⊆ I.
Exercise 4.21. Let R be a PID, and let a, b ∈ R• . Then (a) and (b) can be
factored into products of principal prime ideals, say
(a) = (π1a1 · · · (πiar ), b = (π1b1 ) · · · (πrbr ), ai , bi ∈ N.
min(a ,b )
1 1 min(a ,b )
r r
a) Show ha, bi = hπ1 · · · πr i.
max(a1 −b1 ,0) max(ar −br ,0)
b) Show (a : b) = hπ1 · · · πr i.
c) Show hai ⊂ (a : b).
d) Show (a : b)ha, bi = hai.
e) Suppose #M axSpecR ≥ 2. Find a, b ∈ R• such that:
(i) (a : b) ⊂ ha, bi.
(ii) ha, bi ⊂ (a : b).
(iii) Neither of (a : b), ha, bi contains the other.
We say a partially ordered set (X, ≤) has the weak maximum property if for every
x ∈ X there is a maximal element m ∈ X with x ≤ m. Zorn’s Lemma implies that
if every chain in X has an upper bound then X has the weak maximum property.
5. THE PRIME IDEAL PRINCIPLE OF LAM AND REYES 103

Exercise 4.22. Let F ⊂ I(R) be a family of finitely generated ideals of R.


(This applies to all F iff R is Noetherian.) Show: F 0 has the weak maximum
property.
A family F ⊂ I(R) is an Oka family if R ∈ F and for all x ∈ R and I ∈ I(R), if
hI, xi, (I : x) ∈ F, then I ∈ F.

A family F ⊂ I(R) is an Ako family if R ∈ F and for all x1 , .x2 ∈ R and


I ∈ I(R), if hI, x1 i, hI, x2 i ∈ F, then hI, x1 x2 i ∈ F.

A family F ⊂ I(R) is increasing if for all I, J ∈ I(R), if I ∈ F and J ⊃ I


then J ∈ F.
Theorem 4.27. (Prime Ideal Principle of Lam-Reyes [LR08]) Let R be a ring
and let F ⊂ I(R) be a family of ideals that is either Oka or Ako.
a) The complementary family F 0 := I(R) \ F is an MP family.
b) Suppose moreover that F 0 has the weak maximum property. Then:
(i) Let f ⊂ I(R) be an increasing family. If f ∩ Spec R ⊂ F, then f ⊂ F.
(ii) Suppose that I ∈ I(R) is such that every prime ideal p that contains I
(resp. properly contains I) lies in F. Then every ideal that contains
I (resp. properly contains I) lies in F.
(iii) If Spec R ⊂ F then F = I(R).
Proof. a) We go by contraposition: let I ∈ Max F 0 be an ideal which is not
prime, so there are a, b ∈ R \I with ab ∈ I. Since b ∈ (I : a), the ideals hI, ai, (I : a)
each properly contain I, so by maximality of I we have hI, ai, hI, bi, (I : a) ∈ F.
Since I = hI, abi ∈/ F, the family F is neither Oka nor Ako.
b) (i) Suppose there is I ∈ f \ F. Since F 0 has the weak maximum property, I
is contained in a maximal element p of F 0 , which by part a) is prime. Since f is
increasing, this gives p ∈ f ∩ Spec R ⊂ F, a contradiction.
(ii) Apply (i) with f the family of ideals containing (resp. properly containing) I.
(iii) Apply (i) with f = Spec R. 
Proposition 4.28. Let R be a ring. Each of the following families F is Oka
and the complementary family F 0 is closed under taking unions of chains hence
satisfies the weak maximum property.
(i) The set of all ideals meeting a multiplicatively closed subset S ⊂ R.
(ii) The set of all principal ideals.
(iii) The set of all finitely generated ideals.
Proof. (i) Let x ∈ R, I ∈ I(R) be such that hI, xi, (I : x) ∈ F. Then there
are s1 , s2 ∈ S, i1 , i2 ∈ I and a, b ∈ R such that
s1 = ai1 + bx, s2 x = i2 .
Then
s2 s2 = as2 i1 + bs2 x = as2 i1 + bi2 ∈ S ∩ I.
The union of a chain of ideals disjoint from S is an ideal disjoint from S.
(ii) Suppose (I : x) = hai and hI, xi = hbi. Exercise 4.18d) gives us a useful hint:
we will show I = habi. Equation (15) gives the containment
habi = haihbi(I : x)hI, xi ⊂ I.
104 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Conversely, let i ∈ I. Since I ⊂ hI, xi = hbi, we may write i = αb for some α ∈ R.


We have α(b) ⊂ I hence also αhxi ⊂ I and thus α ∈ (I : x) = hai. So
i = αb ∈ habi.
The union I of a chain {Ij } of nonprincipal ideals must be nonprincipal: if I = hai
then a ∈ Ij for some j and thus hai ⊂ Ij ⊂ I = ha, so Ij = hai is principal.
(iii) Suppose (I : x) = ha1 , . . . , am i and hI, xi = hi1 + α1 x, . . . , in + αn xi. Let
J = hi1 , . . . , in , xa1 , . . . , xam i. We will show I = J, hence I is finitely generated.
It is immediate that J ⊂ I. Conversely z ∈ I; since I ⊂ hI, xi, we may write
z = β1 (i1 + α1 x) + . . . + βn (i1 + αn x) = (β1 i1 + . . . + βn in ) + (α1 β1 + . . . + αn βn )x.
Since z and β1 i1 +. . .+βn in ∈ I, so is (α1 β1 +. . .+αn βn )x, i.e., α1 β1 +. . .+αn βn ∈
(I : x) = ha1 , . . . , am i, so (α1 β1 + . . . + αn βn )x ∈ hxa1 , . . . , xam i and thus z ∈ J.
The union I of a chain {Ij } of infinitely generated ideals must be infinitely
generated: if I = hx1 , . . . , xn i, then for all 1 ≤ i ≤ n we have xi ∈ Iji for some
index ji . Then
hx1 , . . . , xn i ⊂ Imax(j1 ,...,jn ) ⊂ I = hx1 , . . . , xn i,
so Ij = hx1 , . . . , xn i is finitely generated. 
Combining Proposition 4.32 and Theorem 4.27 we deduce a new proof of Multi-
plicative Avoidance as well as immediate proofs of the following results.
Theorem 4.29. If every prime ideal of R is principal, every ideal of R is
principal.
What about maximal ideals? Later we will encounter local rings with maximal
principal ideals that are not principal ideal rings, but they will be non-Noetherian.
Indeed they need to be: by a result of Kaplansky (Theorem 16.9), if in a Noetherian
ring every maximal ideal is principal, then every ideal is principal. The proof will
draw significantly on aspects of the structure theory of Noetherian rings.
Theorem 4.30. (Cohen [Co50]) If every prime ideal of R is finitely generated,
then every ideal of R is finitely generated.
Exercise 4.23. By Exercise 4.18, for any infinite cardinal κ and any ring R,
the family Iκ of ideals of R with fewer than κ generators is an Oka family. So it’s
tempting to generalize Theorem 4.30 to: for any infinite cardinal κ, if every prime
ideal of a ring R can be generated by fewer than κ elements then every ideal of R
can be generated by fewer than κ elements. Is this generalization true?
Looking back at the above approach, the only cloud in the sky may be that directly
checking whether a family is Oka or Ako is neither completely trivial nor espe-
cially enlightening. Following Lam-Reyes, we introduce some further conditions on
a family that will allow us to prove more easily that it is either Oka or Ako.

A family F ⊂ I(R) is a strongly Oka family if R ∈ F and for I, J ∈ I(R),


if hI, Ji, (I : J) ∈ F, then I ∈ F.

A family F ⊂ I(R) is a strongly Ako family if R ∈ F and for all x ∈ R


and I, J ∈ I(R), if hI, xi, hI, Ji ∈ F, then hI, xJi ∈ F.
5. THE PRIME IDEAL PRINCIPLE OF LAM AND REYES 105

A family F ⊂ I(R) is a very strongly Ako family if R ∈ F and for all


I, J1 , J2 ∈ I(R), if hI, J1 i, hI, J2 i ∈ F, then hI, J1 J2 i ∈ F.

A family F ⊂ I(R) is a filter if F is increasing, F is closed under finite inter-


sections and R ∈ F.4 A family F ⊂ I(R) is monoidal if R ∈ F and for all
I, J ∈ F we have IJ ∈ F.
Exercise 4.24. a) Show: a monoidal increasing family F ⊂ I(R) is a
monoidal filter.
b) Let (P) be any of the following properties: Oka, Ako, strongly Oka, strongly
Ako, monoidal filter.
T For each i ∈ I, let Fi be a family satisfying (P).
Show that F := i∈I satsisfies property (P). Deduce: that for any family
F ⊂ I(R) there is a unique minimal family F containing F and satisfying
property (P).
c) Let (P) be the property of being a monoidal filter, and let F ⊂ I(R) be any
family of ideals. Show that F is the collection of ideals of R that contain
a finite product I1 · · · In with each Ii ∈ F.
Proposition 4.31. Let R be a ring, and let F ⊂ I(R) be a family of ideals.
a) Strongly Oka implies Oka, and very strongly Ako implies strongly Ako
implies Ako.
b) Monoidal filter implies very strongly Ako.
c) Very strongly Ako implies strongly Oka.
d) Strongly Ako implies Oka.
Proof. a) These implications are immediate from the definitions.
b) Let I, J1 , J2 ∈ I(R). If hI, J1 i, hI, J2 i ∈ F, then since F is monoidal we have
hI, J1 ihI, J2 i = hI 2 , IJ1 , IJ2 , J1 J2 i ∈ F. Since hI 2 , IJ1 , IJ2 , J1 J2 i ⊂ hI, J1 J2 i and
F is monoidal, we get that hI, J1 J2 i ∈ F, so F is very strongly Ako.
c) Let F be very strongly Ako. Let I, J ∈ I(R) be such that hI, Ji, hI : Ji ∈ F.
Then (I : J) = hI, (I : J)i, and the very strong Ako condition gives
I = hI, J(I : J)i ∈ F.
d) Taking J =]hxi in the proof of part c) shows that strongly Ako implies Oka. 
We now revisit some of the above examples with these further conditions in mind.
Example 4.32. a) If S ⊂ R is a multiplicative subset, the family FS of
all ideals in R that meet S is a monoidal filter, as is almost immediate
to check. This is the strongest of the conditions we have introduced, so in
particular FS is both Oka and Ako.
b) Let F1 be the family of all principal ideals of R. Except in the trivial case
where every ideal of R is principal, F1 is not increasing, so is not a filter.
However it is monoidal and strongly Oka: the former is immediate and the
proof of Oka-ness given in Proposition adapts to show strong Oka-ness.
In fact the family F1 need not be Ako. Let R be a ring admitting
principal ideals hai, hbi whose intersection is not finitely generated (see
[LR08, Remark 3.18]) and the references therein for a specific example).
4This is the usual set-theoretic notion of a filter on 2R each of whose elements is an ideal.
To be slick about it, closure under finite intersections should include closure under the empty
intersection which should itself imply that R ∈ F .
106 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

Then hI, ai = hai, hI, bi = hbi ∈ F1 but hI, abi = I ∈/ F1 . This also gives
a ring in which the family of finitely generated ideals is not Ako (but is
strongly Oka, as we will now see).
c) For an infinite cardinal κ, let F<κ (resp. F≤κ ) be the family of ideals of R
that admit a generating set of cardinality less than α (resp. of cardinality
at most α). These families are monoidal: for F<κ this is because if α, β, κ
are cardinals with κ infinite and α, β < κ then
αβ ≤ max(α, β)2 < κ2 = κ,
and the case of F≤κ is because κ2 = κ. They are also monoidal: we will
treat the slightly more difficult case of F<κ . For an ideal I of R, we write
µ(I) for its minimal number of generators. Let I ∈ I(R) and x ∈ R.
Then there are infinite cardinals α, β < κ such that µ(I, x) ≤ α < κ and
µ((I : x)) ≤ β < κ. Then there is an ideal I0 ⊂ I with µ(I0 ) ≤ α and
hI, xi = hI0 , xi. We claim that
I0 + (I : x)hxi = I.
Indeed, both I0 and (I : x)hxi are contained in I, so I0 + (I : x)hxi ⊂ I. If
i ∈ I there is i0 ∈ I0 and a ∈ R such that i = i0 +ax; since ax = i−i0 ∈ I,
we have a ∈ (I : x), and thus i ∈ I0 + (I : x)hxi. Thus µ(I) ≤ α + β ≤
max(α, β) < κ, so I ∈ Fα .
Exercise 4.25. Let κ be an infinite cardinal, and let F be either F<κ or F≤κ
as defined in Example 4.32c) above. We showed that Max F 0 ⊂ Spec F. Is it true
that if every prime ideal in R can be generated by fewer than κ elements (resp. by
at most κ elements) then every ideal in R can be generated be fewer than κ elements
(resp. by at most κ elements)?

6. Minimal Primes
Let R be a ring. A minimal prime p of R is just what it sounds like: a minimal
element of the set Spec R of prime ideals of R, partially ordered by inclusion.
T
Exercise 4.26. Let C be a chain of prime ideals in a ring R. Show: p∈C p is
a prime ideal.
Proposition 4.33. Let I ⊂ P be ideals of R, with P prime. Then the set S of
all prime ideals p of R with I ⊂ p ⊂ P has a minimal element.
Proof. We partially order S by reverseTinclusion i.e., p1 ≤ p2 ⇐⇒ p1 ⊃ p2 .
Let C be any chain in S. By Exercise 4.26, p∈C p is a prime ideal and thus it is
an upper bound for C in S. By Zorn’s Lemma, S contains a maximal element, i.e.,
a minimal element under ordinary containment. 

Corollary 4.34. Every nonzero ring has at least one minimal prime.
Exercise 4.27. Prove Corollary 4.34.
We write MinSpec R for the set of all minimal primes of R and ZD(R) for the set
of all zerodivisors in R.
T
Exercise 4.28. Show: in a ring R, r(R) = p∈MinSpec R p.
6. MINIMAL PRIMES 107

In order to prove the next result, it is convenient to use the theory of localization,
which we will not develop until § 7. Nevertheless we have decided to place the proof
here, as it fits thematically with the other results of the section.
Theorem S 4.35. Let R be a ring.
a) We have p∈MinSpec R p ⊂ ZD(R).
b) If R is reduced, then equality holds:
[
(16) p = ZD(R).
p∈MinSpec R

Proof. a) Let p ∈ MinSpec R and let x ∈ p. Then pRp is the unique prime
ideal of Rp , so x ∈ r(pAp ) is nilpotent. By Exercise 7.4, this implies that there is
y ∈ R \ p such that yxn = 0. Since y 6= 0, xn – and thus also x – is a zero-divisor.
b) Suppose a ∈ ZD(R), so there is b ∈ R• with ab = 0. Since b 6= 0 and R is
reduced, by Exercise 4.28 we have
\
b∈
/ p,
p∈MinSpec R

so there is p ∈ MinSpec R not containing b. Since ab = 0 ∈ p, we have a ∈ p. 

Proposition 4.36. Let I := {Ij }j∈J be any family of ideals in a ring R, and
let F be the family of ideals that contain some finite product Ij1 · · · Ijn (taking the
empty product, we get that R ∈ F). Let F 0 be the complementary family of ideals
not containing any product of the Ij ’s.
a) Any maximal element of F 0 is a prime ideal.
b) Suppose moreover that each Ii is finitely generated and that every prime
ideal of R contains some Ii . Then there are j1 , . . . , jn ∈ J such that
Ij1 · · · Ijn = (0).
Proof. a) The family F is a monoidal filter: indeed, it is the monoidal filter
generated by the family {Ij }j∈I . By Theorem 4.27a), every maximal element of F 0
is prime.
b) Let {Jk } be a chain in F 0 , and let J := k Jk be its union. If J ⊃ Ij1 · · · Ijn , then
S
since each Iji is finitely generated, so is Ij1 · · · Ijn , and as we have seen before, if
the union of a chain of ideals contains a finitely generated ideal, then so does some
element of the chain. So by Zorn’s Lemma F 0 has the weak maximum property,
so by Theorem 4.27b)(iii) since every prime ideal of R lies in F we conclude that
every ideal lies in F. In particular the zero ideal lies in F, giving the desired
conclusion. 

Corollary 4.37 (Anderson). a) Suppose that in a ring R every mini-


mal prime ideal is finitely generated. Then MinSpec R is finite.
b) If R is Noetherian, then MinSpec R is finite.
Proof. a) We apply Proposition 4.37b) with I = MinSpec R. There are then
distinct minimal prime ideals p1 , . . . , pn and a1 , . . . , an ∈ Z+ such that pa1 1 · · · pann =
(0). If now q ∈ MinSpec R then q ⊃ pa1 1 · · · pann so q ⊃ pi for some 1 ≤ i ≤ n. It
follows that MinSpec R = {p1 , . . . , pn }.
b) If R is Noetherian then every ideal is finitely generated, so part a) applies. 
108 4. FIRST PROPERTIES OF IDEALS IN A COMMUTATIVE RING

7. An application to unit groups


The following useful generalization of Proposition 4.17 is due to MathOverflow user
zcn: see http://mathoverflow.net/users/44201/zcn.
Theorem 4.38. Let f : R → S be a surjective ring homomorphism, with
kernel I. Suppose that all but finitely many maximal ideals of R contain I. Then
the induced group homomorphism on unit groups f × : R× → S × is surjective.
Proof. We may identify S with R/I. Let m1 , . . . , mn be the maximal ideals
of R that do not contain I. Then the ideals I, m1 , . . . , mn are pairwise comaximal.
Let y ∈ S × , and choose x ∈ R such that f (x) = y. By the Chinese Remainder
Theorem there is a ∈ I such that a ≡ 1 − x (mod mi ) for all 1 ≤ i ≤ n. Then
f (x + a) = f (x) + f (a) = f (x) = y. Moreover, for all 1 ≤ i ≤ n we have x + a ∈
/ mi .
If m ∈ MaxSpec R \ {m1 , . . . , mn } then m ⊃ I, so if x + a ∈ m, then x ∈ m, so
y = f (x) ∈ m/I, a proper ideal of R/I: contradiction. So x + a ∈ R× . 
The hypothesis of Theorem 4.38 applies to every surjective homomorphism f : R →
S when R is semilocal, so in particular when R is finite or – as we will see later in
Theorem 8.37 – when R is Artinian.
CHAPTER 5

Examples of Rings

1. Rings of numbers
The most familiar examples of rings are probably rings of numbers, e.g.
Z ⊂ Q ⊂ R ⊂ C.
These are, respectively, the integers, the rational numbers, the real numbers and
the complex numbers. For any positive integer N the ring integers modulo N , de-
noted Z/N Z. We assume that the reader has seen all these rings before.

Historically, the concept of a ring as an abstract structure seems to have arisen


as an attempt formalize common algebraic properties of number rings of various
sorts. It is my understanding that the term “ring” comes from Hilbert’s Zahlring
(“Zahl” means “number” in German). Indeed, various sorts of extension rings of
C – most famously Hamilton’s quaternions H – have been referred to as systems
of hypercomplex numbers. This terminology seems no longer to be widely used.

The adjunction process gives rise to many rings and fields


√ of numbers, as already
seen in §2.2. For instance, for
√ a nonsquare integer D, let D be a complex number
whose square is D: then Z[ D] is an interesting ring.
√ √
Exercise 5.1. Show: Z[ D] = {a + b D | a, b ∈ Z}.

In particular,
√ (Z[ D], +) ∼= (Z2 , +) as commutative groups, although not as rings,
since Z[ D] is a domain and Z2 has nontrivial idempotents.

More generally, let K be any number field (a finite degree field extension of Q),
and let ZK be the set of elements x ∈ K which satisfy a monic polynomial with
Z-coefficients. It turns out that ZK is a ring, the ring of algebraic integers in
K. This is a special case of the theory of integral closure: see §14.

Algebraic number theory proper begins with the observation that in general the
rings ZK need not be UFDs but are otherwise as nice as possible from a commuta-
tive algebraic standpoint. That is, every ring ZK is a Dedekind domain, which
among many other characterizations, means that every nonzero ideal factors into a
product of prime ideals. That the rings ZK are Dedekind domains is an example of
a normalization theorem, more specifically a very special case of the Krull-Akizuki
Theorem of §18.

Let Q be an algebraic closure of Q. (This is not a number field, being an infi-


nite degree algebraic extension of Q.) We may define Z to be the set of all elements
109
110 5. EXAMPLES OF RINGS

of Q which satsify a monic polynomial with integer coefficients: this is the ring of
all algebraic integers. In particular,
Z = lim ZK
−→
is the direct limit of all rings of integers in fixed number fields.
Exercise 5.2. Let Z be the set of all algebraic integers.
a) Taking as given that for any fixed number field K, the algebraic integers in K
form a subring of K, show that Z is a subring of Q.
b) Show: Z is a domain which is not Noetherian. Hint: use the fact that the nth
root of an algebraic integer is an algebraic integer to construct an infinite strictly
ascending chain of principal ideals in Z.
Theorem 5.1. Every finitely generated ideal in the ring Z is principal.
Thus, if only Z were Noetherian, it would be a principal ideal domain! Later on
we will prove a more general theorem, due to Kaplansky, in the context of limits of
Dedekind domains with torsion Picard groups.

2. Rings of continuous functions


2.1. The ring of real-valued functions.

Let R be a ring, X a set, and consider the set RX of all functions f : X → R. We


may endow RX with the structure of a ring by defining addition and multiplication
“pointwise”, i.e.,
(f + g) : x 7→ f (x) + g(x),
(f g) : x 7→ f (x)g(x).
Exercise 5.3. Show: this makes RX into a ring with additive identity the
constant function 0 and multiplicative identity the constant function 1.
However, this is not really a “new” example of a ring.
Exercise 5.4. Show: RX is isomorphic as a ring, to
Q
x∈X R.
Later on we will see this construction in the special case R = F2 , in which case
we get an important subclass of Boolean rings. However, in general RX is quite
a roomy ring. It contains many interesting subrings, some of which can be nicely
consructed and analyzed using topological, geometric and analytic considerations.
2.2. Separation axioms and C(X).

We specialize to the following situation: R = R, X is a topological space, and


instead of the ring RX we look at the subring C(X) of continuous f : X → R.
Exercise 5.5. Show: for a topological space X, the following are equivalent:
(i) For every x, y ∈ X with x 6= y, there exists f ∈ C(X) with f (x) 6= f (y).
(ii) For every x, y ∈ X with x 6= y and every α, β ∈ R, there exists f ∈ C(X) with
f (x) = α, f (y) = β.
(iii) For every finite subset S of X and any function g : S → R, there exists
f ∈ C(X) such that f |S = g.
A space which satisfies these equivalent conditions is called C-separated.1
1More standard terminology: “the continuous functions on X separate points”.
2. RINGS OF CONTINUOUS FUNCTIONS 111

Recall the following chains of implications from general topology:


Lemma 5.2. For any topological space, the following implications hold (and
none of the arrows may be reversed)):
a) X compact =⇒ X normal =⇒ X Tychonoff =⇒ X regular =⇒ X
Hausdorff =⇒ X separated =⇒ X Kolmogorov.
b) X locally compact =⇒ X Tychonoff.
Exercise 5.6. a) Show: a Tychonoff space is C-separated.
b) Show: a C-separated space is Hausdorff.
c)* Show: a regular space need not be C-separated.
(Suggestion: see [Ga71].)
For a topological space X, a zero set is a set of the form f −1 (0) for some continuous
function f : X → R. A cozero set is a complement of a zero set. The cozero sets
in fact form a base for a topology on X; we call it the Z-topology and write XZ
for X endowed with the Z-topology. Since every cozero set is an open set in the
given topology on X, XZ is a coarser topology than the given topology on X. Of
course we allow the possibility that the two topologies coincide. The following basic
(but not so widely known) result gives a condition for this.
Theorem 5.3. a) For a Hausdorff topological space X, the following are equiv-
alent:
(i) XZ = X: every closed set is an intersection of zero sets of continuous functions.
(ii) X is Tychonoff, i.e., if Y is a closed subset of X and x ∈ X \ Y , then there
exists a continuous function f : X → [0, 1] with f (x) = 0, f |Y ≡ 1.
b) For any topological space X, the space XZ is completely regular, and is the finest
completely regular topology on the underlying set of X which is coarser than X.
Proof. [GJ76, p. 38]. 
Let X be a topological space, and let x ∈ X be any point. Consider the set
mx = {f ∈ C(X) | f (x) = 0}.
Evidently mx is an ideal of C(X). But more is true.

Proposition 5.4. Evaluation at x gives a canonical isomorphism C(X)/mx →
R. In particular, mx is a maximal ideal of C(X).
Exercise 5.7. Prove Proposition 5.4.
Thus x 7→ mx gives a map of sets M : X → M (X).
Proposition 5.5. The map M : X → M (X) is injective iff X is C-separated.
Proof. Left to the reader as a routine parsing of the definitions. 
2.3. Quasi-compactness and C(X).
Proposition 5.6. If X is quasi-compact, then M is surjective, i.e., every
maximal ideal of C(X) is of the form mx for at least one point x ∈ X.
Proof. It suffices to show: let I be an ideal of C(X) such that for no x ∈ X
do we have I ⊂ mx . Then I = C(X). By hypothesis, for every x ∈ X there is
fx ∈ I such that fx (x) 6= 0. Since fx is continuous, there is an open neighborhood
Ux of x such fx is nowhere vanishing on Ux . By quasi-compactness of X, there is a
112 5. EXAMPLES OF RINGS

SN
finite set x1 , . . . , xN such X = i=1 Uxi . Then the function f = fx21 + . . . + fx2n is
an element of m which is positive at every x ∈ X. But then f1 is also a continuous
function on X, i.e., f ∈ C(X)× , so I = R. 
A compact space is quasi-compact and C-separated. Thus previous results yield:
Theorem 5.7. If X is compact, then M : X → M (X) is a bijection: every
maximal ideal of C(X) is of the form mx for a unique x ∈ X.
There is a natural topology on M (X), the initial topology: each f ∈ C(X)
induces a function Mf : M (X) → R, by mapping m to the image of f in C(X)/m =
R. We endow M (X) with the coarsest topology which makes each Mf continuous.
Lemma 5.8. For a compact space X, the initial topology on M (X) is Hausdorff.
Proof. For distinct x, x0 ∈ X, consider the maximal ideals mx , mx0 . By C-
separatedness, there exists f ∈ C(X) with f (x) = 0, f (x0 ) 6= 0. Thus choose
disjoint neighborhoods V, V 0 of f (x), f (x0 ) ∈ R. The sets
Uf,V = {x ∈ X | f (x) ∈ V }, Uf,V 0 = {x ∈ X | f (x) ∈ V 0 }
are disjoint open neighborhoods of x and x0 . 
Theorem 5.9. For a compact space X, let M (X) be the set of maximal ideals
of C(X) endowed with the initial topology. Then M : X → M (X), x 7→ mx is a
homeomorphism.
Proof. Step 1: We claim that M is continuous. But indeed, by the universal
property of the initial topology, it is continuous iff for all f ∈ C(X), the composite
function x 7→ mx 7→ f (mx ) is continuous. But this is nothing else than the function
x 7→ f (x), i.e., the continuous function f ! So that was easy.
Step 2: We now know that M is a continuous bijection from a compact space to
a Hausdorff space. Therefore it is a closed map: if Y is closed in X, then Y is
compact, so M(Y ) is compact, so M(Y ) is closed. Therefore M−1 is continuous
and thus M is a homeomorphism. 
2.4. The (Stone-)Zariski topology on C(X).

For any commutative ring R, we define MaxSpec(R) to be the maximal ideals


and put a topology on it: for any ideal I of R, we define
V (I) = {m ∈ MaxSpec R | I ⊂ m}
The sets V (I) are the closed sets for a unique topology on MaxSpec R, the Zariski
topology. Another way to say it is that the closed sets in the Zariski topology are
precisely all sets obtained by intersecting sets of the form
V (f ) = {m ∈ MaxSpec R | f ∈ m}.
To see this, note first that for any ideal I of R,
\
V (I) = V (f )
f ∈I

and for any subset S of R,


\ \
V (f ) = V (f ).
f ∈S f ∈hSiR
2. RINGS OF CONTINUOUS FUNCTIONS 113

The Zariski topology on MaxSpec C(X) was first defined by M. Stone.


Proposition 5.10. Let X be any topological space. The map M : X →
MaxSpec C(X) is continuous when MaxSpec X is given the Zariski topology.
Proof. As above, it is enough to show that for all f ∈ C(X), the preimage
M−1 (V (f )) is closed in X. Unpacking the definitions, we find
M−1 (V (f )) = f −1 (0),
thus the preimage is the zero set of the continuous function f , hence closed. 
Corollary 5.11. For a compact space X, the Zariski topology on M (X) =
MaxSpec C(X) coincides with the initial topology.
Proof. By Theorem 5.9, we may compare the Zariski topology on X – the
topology obtained by pulling back the Zariski topology on MaxSpec C(X) via M
– with the given topology on X. But the proof of Proposition 5.10 shows that the
Zariski topology on X is precisely the Z-topology, i.e., the one in which the closed
subsets are the intersections of zero sets. But X is compact hence quasi-Tychonoff,
so by Theorem 5.3 the Z-topology on X coincides with the given topology on X. 
Now let π : X → Y be a continuous map between compact spaces. There is an
induced map C(π) : C(Y ) → C(X): given g : Y → R, we pullback by π to get
g ◦ π : X → R. It is no problem to see that C(π) is a homomorphism of rings. Now
let mx ∈ MaxSpec C(X) be a maximal ideal and consider its pullback C(π)∗ (mx )
to an ideal of C(Y ): we find
C(π)∗ (mx ) = {g : Y → R | g(π(x)) = 0} = mπ(x) .
Thus the pullback map carries maximal ideals to maximal ideals (recall this is
certainly not true for all homomorphisms of rings!) and thus induces a map from
X to Y which is indeed nothing else than the given map π.
All in all we see that the functors C and MaxSpec give a duality between
the categories of compact spaces and rings of continuous R-valued functions on
compact spaces. In functional analysis this the first step in an important circle of
ideas leading up to Gelfand duality for commutative Banach algebras.
2.5. Further results when X is not compact.
Example 5.12. Let X be an infinite discrete space, so C(X) = RX is the ring of
all functions from X to R. Thus X is a noncompact Tychonoff space. So it follows
from our work so far that M gives a continuous injection from M to the quasi-
compact space MaxSpec C(X). In fact M is an embedding: for any subset Y ⊂ X,
let IY be the ideal of functions vanishing identically on Y . Then the restriction of
the closed subset V (I) of MaxSpec C(X) to M(X) is precisely M(Y ), so M(Y ) is
closed in MaxSpec C(X). Thus M(X) is discrete as a subspace of MaxSpec C(X),
and this implies that M is not surjective.
Theorem 5.13. Let X be any topological space, let M : X → MaxSpec C(X),
and let XT = M(X), viewed as a subspace of MaxSpec C(X).
a) XT is a Tychonoff space.
b) The map M : X → XT is the Tychonoff completion of X: i.e., it is universal
for continuous maps from X to a Tychonoff space.
c) The induced map C(M) : C(XT ) → C(X) is an isomorphism of rings.
114 5. EXAMPLES OF RINGS

Proof. See [GJ76, §3.9]. 

Thus the ring of continuous functions on an arbitrary space X “sees” precisely its
Tychonoff completion XT . Henceforth we restrict to Tychonoff spaces.
Theorem 5.14. Let X be a Tychonoff space. Then the map M : X →
MaxSpec C(X) gives the Stone-Cech compactification of X.
Proof. See [GJ76, §7.11]. 

Corollary 5.15. For any topological space X, MaxSpec C(X) is compact.


Remark 2. We will see later that for any ring R, the Zariski topology on
MaxSpec R is quasi-compact. (In fact this is rather easy, and the interested reader
may well try it now as an exercise.) Thus the content of Corollary 5.15 is that
MaxSpec C(X) is Hausdorff. Later on we will characterize this class of rings and
thereby obtain a different, more elementary proof of Corollary 5.15.
Exercise 5.8. a) Show: C(X) is an R-subalgebra of RX .
b) Show: C(X) is reduced: it contains no nonzero nilpotent elements.
c) (T. Rzepecki) Show: for a topological space, the following are equivalent:
(i) The Tychonoff completion XT of X is a one-point space.
(ii) C(X) = R.
(iii) C(X) is a domain.
(Suggestion: (ii) ⇐⇒ (i) =⇒ (iii) are straightforward. For (iii) =⇒ (ii),
let f ∈ C(X) be nonconstant, so f (x) 6= f (y) for some x, y ∈ X. Show that for
suitable real numbers C1 and C2 the functions g1 = max(0, f1 + C1 ) and g2 =
max(0, −f1 + C2 ) give nonzero elements of C(X) with g1 g2 = 0.)
Exercise 5.9. Show C(X) is connected in the algebraic sense – i.e., there are
no idempotents other than 0 and 1 – iff the topological space X is connected.
Exercise 5.10. Show: there is an antitone Galois connection between 2X and
the set of ideals of C(X), as follows:
S ⊂ X 7→ IS = {f ∈ C(X) | f |S ≡ 0} and
I 7→ YI = {x ∈ X | ∀f ∈ I, f (x) = 0}.
Exercise 5.11. Let X be a Tychonoff space.
a) Let p be a prime ideal of C(X). Show that YI consists of at most one point.
b) Suppose X is moreover compact. Deduce:
(i) YI consists of exactly one point.
(ii) A prime ideal p of C(X) is closed in the sense of the Galois connection – i.e.,
p = IYp iff p is maximal.
(iii) Each prime ideal p of C(X) is contained in a unique maximal ideal.
Exercise 5.12. Let X = [0, 1] with the standard Euclidean topology. Let r0
be the ideal of all functions f ∈ C(X) such that for all k ∈ N, limx→0+ fx(x) k = 0.
Equivalently r0 is the ideal of all functions which are infinitely differentiable at 0
and have identically zero Taylor series at zero.
a) Show: r0 is a radical ideal but not a prime ideal.
b) Show: the only maximal ideal containing r0 is m0 = {f ∈ C([0, 1]) | f (0) = 0}.
c) Deduce that there are ideals of C(X) which are prime but not maximal.
2. RINGS OF CONTINUOUS FUNCTIONS 115

Exercise 5.13. Let X be a C-separated topological space.


a) Let S ⊂ X with #S > 1. Show: IS is not maximal.
b) Suppose X is Tychonoff and S, T ⊂ X. Show: IS ⊂ IT ⇐⇒ T ⊂ S.
c) Show: if X is Tychonoff, then for closed S, T ⊂ X, we have IS = IT ⇐⇒ S = T .
Exercise 5.14. Let ϕ : X → Y be a continuous function between topological
spaces.
a) Show: ϕ induces a ring homomorphism C(ϕ) : C(Y ) → C(X) by g ∈ C(Y ) 7→
ϕ∗ g = g ◦ ϕ.
b) Suppose Y is normal, X is a closed subspace of X and ϕ : X → Y is the inclusion
map. Show: C(ϕ) is surjective.
Exercise 5.15. Let X be a normal topological space. Show: the closure oper-
ator on subsets of X given by the Galois connection is the topological closure.
Exercise 5.16. Let X = {0} ∪ { n1 }n∈Z+ ⊂ R, and let m be the maximal
ideal of functions vanishing at 0. Fill in the details of the following proof that m
is not finitely generated.2 Assume not: m = ha1 , . . . , an i. Then for all g ∈ m,
g 2 (x)
limx→0 |a1 (x)|+...+|a n (x)|
= 0. (Show also that there is δ > 0 such that the denomi-
nator is strictly positive on (0, δ).) Now choose g ∈ m so as to get a contradiction.
Exercise 5.17. Let X be a normal space, and let x ∈ X.
a) Show that the following are equivalent:
(i) The ideal Ix is finitely generated.
(ii) The ideal Ix is principal.
(iii) The point x is isolated in X (i.e., {x} is open).
b) Suppose X is compact. Show that the following are equivalent:
(i) C(X) is a Noetherian ring.
(ii) C(X) is finite-dimensional as an R-vector space.
(iii) X is finite.
Exercise 5.18. Show: if we worked throughout with rings C(X, C) of contin-
uous C-valued functions, then all of the above results continue to hold.
Exercise 5.19. Suppose we looked at rings of continuous functions from a
topological space X to Qp . To what extent to the results of the section continue to
hold?
Exercise 5.20. Let X be a compact smooth manifold and consider the ring
C ∞ (X) of smooth functions f : X → R.
a) Show: for x ∈ X, {f ∈ C ∞ (X) | f (x) = 0} is a finitely generated maximal
ideal.
b) The phenomenon of part a) is in contrast to the case of maximal ideals in the
ring C([0, 1]), say. However, I believe that with this sole exception, all of the results
of this section hold for the rings C ∞ (X) just as for the rings C(X). Try it and see.
2.6. A theorem of B. Sury.
Theorem 5.16. (Sury) Let c ∈ [0, 1], and let mc = {f ∈ C[0, 1] | f (c) = 0}.
Then mc admits no countable generating set.

2Or, if you like, give your own proof that m is not finitely generated!
116 5. EXAMPLES OF RINGS

Proof. Let {fn }∞ ∞


n=1 be a countably infinite subset of mc , and let J = h{fn }n=1 i.
It suffices to exhibit fT∈ mc \ J. By rescaling, we may assume ||fn || ≤ 1 for all n.

And we may assume n=1 fn−1 (0) = {c}: otherwise x 7→ |x − c| lies in mc \ J. Put

r
X |fn (x)|
f (x) = .
n=1
2n

The series is uniformly convergent (by “Weierstrass’s M-Test”) and thus f , being
the uniform limit of continuous functions, is itself continuous. Moreover f −1 (0) =
{c}, and in particular f ∈ mc . Seeking a contradiction, we suppose f ∈ J: then
there is r ∈ Z+ and g1 , . . . , gr ∈ C([0, 1]) such that
r
X
f= gn fn .
n=1
Pr
Let M = max √ 1≤n≤r ||gn ||, so ||f || ≤ M n=1 ||fn ||. P
Let U be a neighborhood of c
r
such that || fn ||U < 2n1M for 1 ≤ n ≤ r. Since f = n=1 gn fn vanishes only at c,
for each x ∈ U \ {c}, there exists 1 ≤ N ≤ r such that fN (x) 6= 0 and thus
p
|fN (x)|
|fN (x)| < .
2N M
Hence
r r p
X X |fn (x)|
|f (x)| ≤ M |fn (x)| < ≤ |f (x)|,
n=1 n=1
2n
a contradiction. 

3. Rings of holomorphic functions


We have just seen that the ring of continuous functions on a topological space is
very rarely a domain. A remedy for this is to consider more “rigid” collections of
functions. Let U be an open subset of the complex plane C, and let Hol(U ) be
the set of holomorphic functions f : U → C. (A holomorphic function on U is
one for which the complex derivative f 0 (z) exists for each z ∈ U . Equivalently, for
each z ∈ U f admits a power series expansion with positive radius of convergence.)
Then Hol(U ) ⊂ CU , the ring of all C-valued functions on U .
Proposition 5.17. For nonempty open U ⊂ C, the following are equivalent:
(i) U is connected.
(ii) Hol(U ) is a domain.
Proof. (i) =⇒ (ii): For any f ∈ C(U, C) let Z(f ) = {z ∈ U | f (z) = 0} be
the zero set of f . Since f is continuous, Z(f ) is a closed subset of U . If f is moreover
holomorphic, then Z(f ) has no accumulation point in U , i.e., f 6= 0 =⇒ Z(f )
is discrete – in particular Z(f ) is countable. Moreover, for any f, g ∈ C(U, C) we
have Z(f g) = Z(f ) ∪ Z(g), so if f, g ∈ Hol(U )• then Z(f g) is at most countable,
whereas U is uncountable, so f g 6= 0.
¬ (i) =⇒ ¬ (ii): If U is not connected, then U = V1 ∪V2 where V1 and V2 are disjoint
open subsets. Let χi be the characteristic function of Vi for i = 1, 2. Then each χi
is locally constant on U – hence holomorphic, and nonzero, but χ1 χ2 = 0. 
3. RINGS OF HOLOMORPHIC FUNCTIONS 117

In complex function theory it is common to call a nonempty connected open U ⊂ C


a domain. Henceforth we assume that U is a domain. For z ∈ U there is a function
ordz : Hol(U )• → N,Pthe order of vanishing of f at z: we expand f into a power

series at z: f (ζ) = n=0 an (ζ −z)n and let ordz (f ) be the least n for which an 6= 0.
Compiling these we associate to each f ∈ Hol(U )• its total order Ord(f ) : U → N
given by Ord(f )(z) = ordz (f ). Consider the set NU of all functions from U to N.
For O ∈ NU , we define the support of O to be the set of z ∈ U such that O(z) > 0.

A meromorphic function on U is a function which is holomorphic on U except


for isolated finite order singularities. More precisely, a meromorphic function is a
function which is holomorphic on U \ Z for some discrete closed subset Z of U and
such that for all z0 ∈ Z, there exists n ∈ Z+ such that (z − z0 )n f (z) extends to a
holomorphic function on a neighborhood of z. If the least n as above is positive,
we say that f has a pole at z0 , and we employ the convention that f (z0 ) = ∞. Let
Mer(U ) be the set of all meromorphic functions on U ; it is a ring under pointwise
addition and multiplication, under the conventions that for all z ∈ C,
z + ∞ = ∞ + ∞ = z · ∞ = ∞ · ∞ = ∞.
Theorem 5.18. (Weierstrass + Mittag-Leffler) Let U ⊂ C be a domain.
a) If O ∈ NU has closed, discrete support, there is f ∈ Hol(U )• with Ord(f ) = O.
b) Let Z ⊂ U be a closed subset without limit points. To each z ∈ Z we associate
nz ∈ N and wz,k ∈ C for all 0 ≤ k ≤ nz . Then there is f ∈ Hol(U ) such that for
all z ∈ Z and 0 ≤ k ≤ nz , f (k) (z) = k!wz,k .
Proof. Part a) is part of Weierstrass’ Factorization Theory [Ru87, Thm.
15.11]. To get part b), combine part a) with Mittag-Leffler’s result on the existence
of meromorphic functions with prescribed principal parts [Ru87, Thm. 15.13]. 
Corollary 5.19. The ring Mer(U ) of meromorphic functions on U is a field,
and indeed is the field of fractions of Hol(U ).
Exercise 5.21. Prove Proposition 5.19.
Exercise 5.22. Fix z0 ∈ U . For f ∈ Mer(U ), choose n ∈ N such that (z−z0 )n f
is holomorphic at z0 , and put ordz0 (f ) = ordz0 ((z − z0 )n f ) − n.
a) Show that this gives a well-defined function ordz0 : Mer(U )• → Z (i.e., indepen-
dent of the choice of n in the definition).
b) Show that for all f, g ∈ Mer(U )× , ordz0 (f g) = ordz0 (f ) + ordz0 (g).
c) We formally extend ordz0 to a function from Mer(U ) to Z ∪ {∞} by setting
ordz0 (0) = ∞. Show that, under the convention that ∞ + n = ∞ + ∞ = ∞, we
have for all f, g ∈ Mer(U ) that ordz0 (f + g) ≥ min ordz0 (f ), ordz0 (g).
d) Show that if ordz0 (f ) 6= ordz0 (g) then ordz0 (f + g) = min ordz0 f, ordz0 (g).
Similarly we may extend Ord to a function from Mer(U )• to ZU .
Lemma 5.20. For f, g ∈ Hol(U )• , the following are equivalent:
(i) Ord(f ) = Ord(g).
(ii) f = ug for u ∈ Hol(U )× .
(iii) (f ) = (g).
Proof. (ii) ⇐⇒ (iii) for elements of any domain.
(ii) =⇒ (i) is easy and left to the reader.
118 5. EXAMPLES OF RINGS

(i) =⇒ (ii): The meromorphic function fg has identically zero order, hence is
nowhere vanishing and is thus a unit u in Hol(U ). 
Theorem 5.21. (Helmer [He40]) For a domain U in the complex plane, every
finitely generated ideal of Hol(U ) is principal. More precisely, for any f1 , . . . , fn ∈
Hol(U )• , there exists f ∈ Hol(U ) such that Ord(f ) = mini Ord(fi ), unique up to
associates, and then hf1 , . . . , fn i = hf i.
Proof. Step 1: Suppose f1 , f2 ∈ Hol(U )• don’t both vanish at any z ∈ U . Let
Z be the zero set of f1 , so for all z ∈ Z, f2 (z) 6= 0. Theorem 5.18b) gives g2 ∈ Hol(U )
such that for all z ∈ Z, ordz (1 − g2 f2 ) ≥ ordz (f1 ). Thus Ord(1 − g2 f2 ) ≥ Ord(f1 ),
so g1 := 1−gf1
2 f2
∈ Hol(U ), f1 g1 + f2 g2 = 1 and hf1 , f2 i = Hol(U ).
Step 2: Now let f1 , f2 ∈ Hol(U )• be arbitrary. By Theorem 5.18a), there exists
f ∈ Hol(U ) with Ord(f ) = min Ord(f1 ), Ord(f2 ). For i = 1, 2, put gi = ffi . Then g1
and g2 are holomorphic and without a common zero, so by Step 1 hg1 , g2 i = Hol(U ).
Multiplying through by f gives hf1 , f2 i = hf i.
Step 3: If in a ring every ideal of the form hx1 , x2 i is principal, then every finitely
generated ideal is principal. By Step 2, this applies in particular to Hol(U ). More-
over, if the ideal hf1 , . . . , fn i = hf i, then we must have Ord f = min Ord fi . 
Exercise 5.23. Explain why in Step 1 above, Theorem 5.18b) implies that g2
exists.
The most familiar domains in which every finitely generated ideal is principal are
those in which every ideal is principal: PIDs! But as the reader has probably al-
ready suspected, Hol(U ) is not a PID. One way to see this is to show that Hol(U )
is not even a UFD. Remarkably, this is a consequence of the Weierstrass Factor-
ization Theory, which expresses every holomorphic function as a product of prime
elements! The catch is that most holomorphic functions require infinite products,
a phenomenon which is not countenanced in the algebraic theory of factorization.
Exercise 5.24. Let f ∈ Hol(U )• .
a) Show that f is an irreducible element of Hol(U ) – i.e., if f = g1 g2 then exactly
one of g1 , g2 is a unit – iff it has exactly one simple zero.
b) Suppose f is irreducible. Show Hol(U )/(f ) = C. In particular, (f ) is prime.
c) Show that f admits a (finite!) factorization into irreducible elements iff it has
only finitely many zeros. Conclude that Hol(U ) is not a UFD.
Exercise 5.25. a) Show: all the results of this section extend to the ring of
holomorphic functions on a noncompact Riemann surface.
b) Investigate which of the results of this section hold for all Stein manifolds.3

4. Kapovich’s Theorem
4.1. The cardinal Krull dimension of a partially ordered set.

Throughout, all rings are commutative and with multiplicative identity. For a
ring R, Spec R is the set of prime ideals of R, partially ordered under inclusion.

A chain is a linearly ordered set; its length is its cardinality minus one. The
3Step 1: learn the definition of a Stein manifold!
4. KAPOVICH’S THEOREM 119

cardinal Krull dimension carddim X of a partially ordered set X is the supre-


mum of lengths of its chains. For a ring R we put carddim R = carddim Spec R.
Remark 3. The prime spectrum Spec R of a ring is endowed with the Zariski
topology, in which the closed sets are V (I) = {p ∈ Spec R | p ⊃ I} as I ranges
over all ideals of R. For p1 , p2 ∈ Spec R we have p1 ⊂ p2 ⇐⇒ p2 ∈ {p1 }. Thus
carddim R is a topological invariant of Spec R.
For a topological space X, define the cardinal Krull dimension carddim X as
the supremum of lengths of chains of closed irreducible subspaces of X. Since for
a ring R the map p 7→ V (p) gives an antitone bijection from Spec R to the set of
closed irreducible subspaces of Spec R, we have carddim R = carddim Spec R.
Our use of the word “cardinal” is twofold: (i) it is common to say “dim R is infinite”
if there are arbitrarily long finite chains in Spec R. For the class of rings considered
here we will show the Krull dimension is zero or infinite, but we will not completely
answer the more refined question of how infinite it is. (ii) There is also a notion of
ordinal Krull dimension of rings [GoRo] that we do not discuss here.
Remark 4.
a) Let X and Y bre partially ordered sets. If there is an injective isotone map
ι : Y → X, then carddim Y ≤ carddim X.
b) If f : R1 → R2 is surjective or a localization map, then f ∗ : Spec R2 → Spec R1
is an injective isotone map, so carddim R2 ≤ carddim R1 .
4.2. Holomorphic functions on a C-manifold.

Let M be a C-manifold. (Our definition includes that M is Hausdorff and second


countable.) Let Hol(M ) be the ring of global holomorphic functions f : M → C.
We have C ,→ Hol(M ) via the constant functions.
Lemma 5.22. The ring Hol(M ) is a domain iff M is connected.
`
Proof. If M = M1 M2 with M1 , M2 6= ∅, let fi be the characteristic func-
tion of Mi . Then f1 , f2 ∈ Hol(M )• and f1 f2 = 0.
Conversely, let f ∈ Hol(M )• , and let U be the set of x ∈ M such that the
power series expansion at x is zero (as a formal series: i.e., every term is zero).
For all x ∈ U , f vanishes identically in some neighborhood of x, so U is open. If
x ∈ M \ U , then some mixed partial derivative of f is nonvanishing at x. These
mixed partials are continuous, so there is a neighborhood Nx of x on which this
condition continues to hold, and thus Nx ⊂ M \ U and U is closed. Since M is
connected and U ( M , we have U = ∅. For f, g ∈ Hol(M )• , let x ∈ M . The power
series of f and g at x are each nonzero, hence the same holds for f g. So f g does
not vanish identically on any neighborhood of x: thus f g 6= 0. 
From now on we will assume that all our C-manifolds are connected.
4.3. Kapovich’s Theorems: Statements.
Theorem 5.23. (Kapovich) Let M be a C-manifold. Then either Hol(M ) = C
or carddim Hol(M ) ≥ c = 2ℵ0 .
A discrete valuation on a ring R is a surjective function
v : R → N ∪ {∞}
120 5. EXAMPLES OF RINGS

such that
(DV0) For all x ∈ R, v(x) = ∞ ⇐⇒ x = 0.
(DV1) For all x, y ∈ R, v(xy) = v(x) + v(y).
(DV2) For all x, y ∈ R, v(x + y) ≥ min v(x), v(y).

Here we use some standard conventions on arithmetic in the extended real numbers:
for all x ∈ [0, ∞], x + ∞ = ∞ and min(x, ∞) = x. Conditions (DV0) and (DV1)
ensure that a ring that admitting a discrete valuation is a domain.

A V∞ -ring is a ring R admitting a sequence {vk }k∈Z+ of discrete valuations such


that for any sequence {nk }∞ •
k=1 of natural numbers there is x ∈ R such that
+
vk (x) = nk for all k ∈ Z . The following results together imply Theorem 5.23.
Theorem 5.24. If R is a V∞ -ring, then carddim R ≥ c.
Theorem 5.25. Let M be a C-manifold. If M admits a nonconstant holomor-
phic function, then Hol(M ) is a V∞ -ring.
4.4. Preliminaries on ultralimits.

Let I be a set, let X be a topological space, and let x• : I → X be a func-


tion. Let F be an ultrafilter on I. We say x ∈ X is an ultralimit of x• and
write F lim x• = x if x• (F) → x: that is, for every neighborhood U of x ∈ X,
we have x−1 • (U ) ∈ F. From the general theory of filter convergence, we deduce:
(i) If X is Hausdorff, then every I-indexed sequence x• : I → X has at most one
ultralimit. (ii) If X is quasi-compact, then every I-indexed sequence has at least
one ultralimit. Thus (iii) If X is compact, then every I-indexed sequence has a
unique ultralimit. In our application we will have I = N, ω a fixed nonprincipal
ultrafilter and X = [0, ∞]. Thus we have an ordinary sequence {xn } in [0, ∞], and
ω lim xn = x means: for all  > 0, the set of n ∈ N such that |xn − x| <  lies in ω.
Because [0, ∞] is compact, any sequence in [0, ∞] has a unique ultralimit.
Remark 5. Let ω be a nonprincipal ultrafilter on Z+ .
a) Show: if limk→∞ xk = x in the usual sense, then also ω limk xk = x.
b) Let {xk }, {yk } be sequences in [0, ∞]. Show:
(i) ω limk (xk + yk ) = ω limk xk + ω limk yk .
(ii) ω limk min(xk , yk ) = min(ω limk xk , ω limk yk ).
(iii) ω limk max(xk , yk ) = max(ω limk xk , ω limk yk ).
4.5. Proof of Theorem 5.24.

For t ∈ (0, ∞), put


vk (x)
pt = {x ∈ R• | ω lim > 0}.
k kt
Each pt is a prime ideal, and for all t1 ≥ t2 we have pt1 ⊂ pt2 . Since R is a V∞ -ring,
there is xt ∈ R• such that vk (xt ) = dk t e for all k ∈ Z+ , and we have xt ∈ pt , xt ∈
/ ps
for all s > t. So {pt | t ∈ (0, ∞)} is a chain of prime ideals of R of cardinality c.
4.6. Proof of Theorem 5.25.

Let h : M → C be holomorphic and nonconstant. By the Open Mapping The-


orem, U = h(M ) is a connected open subset of C. In particular U is metrizable
4. KAPOVICH’S THEOREM 121

and not compact, so there is a sequence {zk }∞ k=1 of distinct points of U with no
accumulation point in U . We do not disturb the latter property by successively
replacing each zk with any point in a sufficiently small open ball, so by Sard’s
Theorem we may assume that each zk is a regular value of h. For k ∈ Z+ , let
pk ∈ h−1 (zk ) and let vk : Hol(M )• → N be the order of vanishing of h at pk : that
is, the least N such that there is a mixed partial derivative of order N which is
nonvanishing at pk . Then vk is a discrete valuation. Let {nk }∞k=1 be as sequence of
natural numbers. By the Weierstrass Factorization Theorem [Ru87, Thm. 15.11],
there is g ∈ Hol(U ) such that ordzk (g) = nk and thus – since pk is a regular value
for h – for all k ∈ Z+ we have vk (g ◦ h) = nk .
4.7. The cardinal Krull dimension of a Stein manifold.

We will now prove a stronger lower bound on the cardinal Krull dimension of
Hol(M ) for when M is a Stein manifold: a C-manifold which admits a closed
(equivalently proper) holomorphic embedding into CN for some N ∈ Z+ . Stein
manifolds play the role in the biholomorphic category that affine varieties play in
the algebraic category (of quasi-projective varieties V/C , say) – and a nonsingu-
lar affine variety over C is a Stein manifold – namely the C-manifolds which have
“enough” global holomorphic functions: in particular, for points x 6= y on a Stein
manifold M , there is f ∈ Hol(M ) with f (x) 6= f (y). At the other extreme lie the
compact C-manifolds, which play the role in the biholomorphic category that pro-
jective varieties play in the algebraic category – and a nonsingular projective variety
over C is a compact C-manifold). In dimension one this is a simple dichotomy: a
Riemann surface is a Stein manifold iff it is noncompact [GuRo, p. 209].
Theorem 5.26. If S1 , S2 are noncompact Riemann surfaces then
carddim Hol(S1 ) = carddim Hol(S2 ) ≥ 2ℵ1 .
Proof. Henriksen showed Hol(C) ≥ 2ℵ1 [He53]. For noncompact Riemann
surfaces S and T , Alling showed Spec Hol(S) and Spec Hol(T ) are homeomorphic
[Al63]. By Remark 3 it follows that carddim Hol(S) = carddim Hol(C) ≥ 2ℵ1 . 
Lemma 5.27. Let M1 , M2 be C-manifolds. Then
carddim Hol(M1 × M2 ) ≥ carddim Hol(M1 ).
Proof. Fix y0 ∈ M2 . Pulling back holomorphic functions via the embedding
ι : M1 ,→ M1 × M2 , x 7→ (x, y0 )
gives a ring homomorphism ι∗ : Hol(M1 × M2 ) → Hol(M1 ). If f ∈ Hol(M1 ) put
F : M1 × M2 → C, (x, y) 7→ f (x).
Then F ∈ Hol(M1 × M2 ) and ι∗ (F ) = f . So we may apply Remark 4b). 
Theorem 5.28. Let M be a C-manifold of the form V ×N for a Stein manifold
V . Then carddim Hol(M ) ≥ carddim Hol(C) ≥ 2ℵ1 .
Proof. Lemma 5.27 reduces us to the case in which M is a Stein mani-
fold. If f : M → C is a nonconstant holomorphic function, then a connected
component M 0 of the preimage of a regular value is a closed submanifold with
dimC M 0 = dimC M − 1. A closed C-submanifold of a Stein manifold is a Stein
manifold [GuRo, p. 210], so we may repeat the process, eventually obtaining a
122 5. EXAMPLES OF RINGS

closed embedding ι : S ,→ M with S a connected, one-dimensional Stein mani-


fold, hence a connected, noncompact Riemann surface. Now if Y is a closed C-
submanifold of a Stein manifold X then the map Hol X → Hol Y obtained by
restricting holomorphic functions to Y is surjective [GuRo, Thm. VIII.18], so
ι∗ : Hol(M ) → Hol(S) is surjective. By Remark 4b) and Theorem 5.26 we have
carddim M ≥ carddim S ≥ 2ℵ1 . 
4.8. Final Remarks.

A little set theory: For a C-manifold M , the ring Hol(M ) is a subring of the
ring of all continuous C-valued functions. For any separable topological space X,
the set of continuous functions f : X → C has cardinality at most cℵ0 = (2ℵ0 )ℵ0 =
2ℵ0 ×ℵ0 = 2ℵ0 = c. Since C ⊂ Hol(M ) we have # Hol(M ) = c. It follows that
Hol(M ) has at most 2c ideals and thus carddim Hol(M ) ≤ 2c . Moreover
c = 2ℵ0 ≤ 2ℵ1 ≤ 2c .
Whether either inequality is strict is independent of the ZFC axioms, but e.g.
the Continuum Hypothesis (CH) gives c < 2ℵ1 = 2c . Thus under CH we have
carddim Hol(M ) = 2c for any Stein manifold M . It may well be the case that the
determination of carddim Hol(C) is independent of the ZFC axioms.

A little history: Theorem 5.23 is a result of M. Kapovich. It answers a question


of G. Elencwajg: is there a C-manifold M with carddim Hol(M ) finite and positive?
Kapovich’s construction draws from work of Henriksen [He53] and Sasane [Sa08].
Kapovich’s first proof of Theorem 5.24 was closely modelled on a criterion of Sasane
[Sa08, Thm. 2.2] which was used by Sasane to show that the Krull dimension of
a noncompact Riemann surface is infinite (which had earlier been established by
Alling). I believe Sasane’s proof is faulty: [Sa08, (2.6)] assumes that the Multi-
plicative Avoidance Theorem gives a unique ideal. I corresponded with Kapovich,
and he immediately replaced the limsup with an ultralimit.
The proof of Theorem 5.23 given here is directly inspired by Kapovich’s proof.
The proof of Theorem 5.25 is identical to Kapovich’s, but the proof of Theorem 5.24
is a bit different. Kapovich uses hyperreals and hypernaturals, but in an earlier ver-
sion he used ultralimits to show carddim Hol(M ) ≥ 1 =⇒ carddim Hol(M ) ≥ ℵ0 ;
we adapt this argument to show the stronger result by following a construction
of Henriksen [He53]. Henriksen’s proof is not couched in the language of ultra-
limits, but this is just an expository difference: crucially, it uses a nonprincipal
ultrafilter on Z+ . Our perspective is that ultrafilters are a nice way to package this
argument, as it makes the bookkeeping virtually automatic. Kapovich uses Multi-
plicative Avoidance, a result which is equivalent in ZF Set Theory to the Boolean
Prime Ideal Theorem. The existence of an ultrafilter on Z+ is a somewhat weaker
set-theoretic axiom but still not a theorem in ZF. I wonder (idly and inexpertly)
whether carddim Hol(C) ≥ ℵ0 can be proved in ZF set theory alone.

5. Polynomial rings
Let R be a ring (possibly non-commutative, but – as ever – with identity). Then
R[t] denotes the ring of univariate polynomials with R-coefficients.

We assume the reader knows what this means in at least an informal sense: an
5. POLYNOMIAL RINGS 123

element of R will be an expression of the form an tn + . . . + a1 t + a0 , where n is some


non-negative integer and an , . . . , a0 are in R. The degree of a polynomial is the
supremum over all numbers n such that an 6= 0. We say “supremum” rather than
“maximum” as an attempt to justify the convention that the degree of the 0 poly-
nomial should be −∞ (for that is the supremum of the empty set). A polynomial
of degree 0 is called constant, and we can view R as a subset of R[t] by mapping
a ∈ R to the constant
L∞ polynomial a. As an commutative group, R[t] is canonically
tn 7→ (a0 , a1 , . . .).
P
isomorphic to n=0 R, the isomorphism being given by n an
(The key point here is that on both sides we have an = 0 for all sufficiently large
n.) Multiplication of polynomials is obtained by applying the relations t0 = 1,
ti+j = ti tj at = ta for all a ∈ R, and distributivity, i.e.,
X
(an tn + . . . + an t1 + a0 ) · (bm tm + . . . + b1 t + b0 ) = ai bj ti+j .
0≤i≤n, 0≤j≤m

For any P ∈ R[t], the identity 1 ∈ R has the property 1 · P = P · 1 = 1.

Unfortunately there are some minor annoyances of rigor in the previous descrip-
tion. The first one – which a sufficiently experienced reader will immediately either
dismiss as silly or know how to correct – is that it is not set-theoretically correct:
technically speaking, we need to say what R[t] is as a set and this involves saying
what t “really is.” It is common in abstract algebra to refer to t is an indeter-
minate, a practice which is remarkably useful despite being formally meaningless:
essentially it means “Don’t worry about what t is; it can be anything which is not
an element of R. All we need to know about t is encapsulated in the multiplication
rules at = ta, t0 = 1, ti tj = ti+j .” In other words, t is what in the uncomplicated
days of high school algebra was referred to as a variable.

If someone insists that R[t] be some particular set – a rather unenlighened atti-
tude that we will further combat
L∞ later on – then the solution has already been
given: we can take R[t] = n=0 R. (It is fair to assume that we already know
what direct sums of commutative groups “really are”, but in the next section we
will give a particular construction which is in fact rather useful.) This disposes of
the set-theoretic objections.

Not to be laughed away completely is the following point: we said R[t] was a
ring, but how do we know this? We did explain the group structure, defined a mul-
tiplication operation, and identified a multiplicative identity. It remains to verify
the distributivity of multiplication over addition (special cases of which motivated
our definition of multiplication, but nevertheless needs to be checked in general)
and also the associativity of multiplication.

Neither of these properties are at all difficult to verify. In fact:

Exercise 5.26. a) Show: R[t] is a ring.


b) Show: R[t] is commutative iff R is commutative.

Let us now attempt a “conceptual proof” of the associativity of polynomial multi-


plication. For this we shall assume that R is commutative – this is the only case
we will be exploring further anyway. Then we can, as the P (t) notation suggests,
124 5. EXAMPLES OF RINGS

view an element of R[t] as a function from R to R. Namely, we just plug in values:


a ∈ R 7→ P (a) ∈ R.
To be clear about things, let us denote this associated function from R to R by
P . As we saw above, the set of all functions RR from R to R forms a commuta-
tive ring under pointwise addition and multiplication: (f + g)(a) := f (a) + g(a),
(f g)(a) := f (a) · g(a). In particular, it really is obvious that the multiplication of
functions is associative. Let P be the subset of RR of functions of the form P for
some P ∈ R[t]. More concretely, we are mapping the constant elements of R[t] to
constant functions and mapping t to the identity function. This makes it clear that
P is a subring of RR : in fact it is the subring of RR generated by the constant
functions and the identity function.

So why don’t we just define R[t] to be P, i.e., identify a polynomial with its asso-
ciated function?

The problem is that the map R[t] → P need not be an injection. Indeed, if R
is finite (but not the zero ring), P is a subring of the finite ring RR so is obviously
finite, whereas R[t] is just as obviously infinite. If R is a domain this turns out to
be the only restriction.
Proposition 5.29. Let R be a domain.
a) Suppose that R is infinite. Then the canonical mapping R[t] → P is a bijection.
b) Suppose that R is finite, say of order q, and is therefore a field. Then the kernel
of the canonical mapping R[t] → P is the principal ideal generated by tq − t.
We leave the proof as a (nontrivial) exercise for the interested reader.
Exercise 5.27. Exhibit an infinite commutative ring R for which the map
R[t] → P is not injective. (Suggestion: find an infinite ring all of whose elements
x satisfy x2 = x.)
Exercise 5.28. Show: the map R[t] → P is a homomorphism of rings.
So if we restrict to infinite domains, the map R[t] → P is an isomorphism of rings.
Thus we see, after the fact, that we could have defined the ring structure in terms
of pointwise multiplication.

6. Semigroup algebras
A semigroup M is a set equipped with a single binary operation ·, which is required
(only!) to be associative. A monoid is a semigroup with a two-sided identity.
Exercise 5.29. Show: a semigroup has at most one two-sided identity, so it
is unambiguous to speak of “the” identity element in a monoid. We will denote it
by e (so as not to favor either addditive or multiplicative notation).
Example 5.30. Let (R, +, ·) be an algebra. Then (R, ·) is a semigroup. If R is
a ring (i.e., has an identity 1) then (R, ·) is a monoid, with identity element 1.
Example 5.31. Any group is a monoid. In fact a group is precisely a monoid
in which each element has a two-sided inverse.
Example 5.32. The structure (N, +) of natural numbers under addition is a
monoid; the identity element is 0.
6. SEMIGROUP ALGEBRAS 125

Example 5.33. The structure (Z+ , ·) of positive integers under multiplication


is a monoid; the identity element is 1.
Let M and N be two semigroups. Then the Cartesian product M × N becomes
a semigroup in an obvious way: (m1 , n1 ) · (m2 , n2 ) := (m1 · m2 , n1 · n2 ). If M
and N are monoids with identity elements eM and eN , then M × N is a monoid,
with identity element (eM , eN ). Exactly the same discussion holdsL for any finite
n
set M1 , . . . , MN of semigroups: we can form the direct sum M = i=1 Mi , i.e.,
the Cartesian product of sets with componentwise operations; if all the Mi ’s are
monoids, so is M .
If we instead have an infinite family {Mi }i∈I of semigroups Q
indexed by a set I,
we can define a semigroup structure on the Cartesian product i∈I Mi in the ob-
vious way, and if each Mi is a monoid with identity ei , then the product semigroup
is a monoid with L identity (ei )i∈I . If each Mi is a monoid, we Q
can also define the
direct sum i∈I Mi , which is the subset of the direct product i∈I Mi consisting
L (mi ∈ Mi )i∈I such that mi = ei for all but finitely many
of all I-tuples Q i. Then we
have that i∈I Mi is a submonoid of the direct product monoid i∈I Mi .

If M and N are semigroups, then a map f : M → N is a homomorphism of


semigroups if f (m1 · m2 ) = f (m1 ) · f (m2 ) for all m1 , m2 ∈ M . If M and N are
monoids, a homomorphism of monoids is a homomorphism of semigroups such that
moreover f (eM ) = eN . A homomorphism f : M → N of semigroups (resp. of
monoids) is an isomorphism iff there exists a homomorphism of semigroups (resp.
monoids) g : N → M such that g ◦ f = IdM , f ◦ g = IdN .
Exercise 5.30. a) Exhibit monoids M and N and a homomorphism of semi-
groups f : M → N which is not a homomorphism of monoids.
b) Show that a homomorphism of semigroups f : M → N is an isomorphism iff it
is bijective. Same for monoids.
+
Exercise 5.31. Show:L∞the monoid (Z , ·) of positive integers under multipli-
cation is isomorphic to i=1 (N, +), i.e., the direct sum of infinitely many copies
of the natural numbers under addition. (Hint: a more natural indexing set for the
direct sum is the set of all prime numbers.)
Now let R be an algebra and M be a semigroup. We suppose first that M is finite.
Denote by R[M ] the set of all functions f : M → R.

As we saw, using the operations of pointwise addition and multiplication endow


this set with the structure of an associative algebra (which has an identity iff M
does). We are going to keep the pointwise addition but take a different binary
operation ∗ : R[M ] × R[M ] → R[M ].

Namely, for f, g ∈ R[M ], we define the convolution product f ∗ g as follows:


X
(f ∗ g)(m) := f (a)g(b).
(a,b)∈M 2 | ab=m

In other words, the sum extends over all ordered pairs (a, b) of elements of M whose
product (in M , of course), is m.
Proposition 5.34. Let R be an associative algebra and M a finite semigroup.
The structure (R[M ], +, ∗) whose underlying set is the set of all functions from M to
126 5. EXAMPLES OF RINGS

R, and endowed with the binary operations of pointwise additition and convolution
product, is an associative algebra. If R is a ring and M is a monoid with identity
e, then R[M ] is a ring with multiplicative identity the function I which takes eM
to 1R and every other element of M to 0R .
Proof. First, suppose that R is a ring and M is a monoid, then for any
f ∈ R[M ] and m ∈ M , we have
X
(f ∗I)(m) = f (a)I(b) = f (m)I(1) = f (m) = I(1)f (m) = . . . = (I∗f )(m).
(a,b)∈M 2 | ab=m

We still need to check the associativity of the convolution product and the distribu-
tivity of convolution over addition. We leave the latter to the reader but check the
former: if f, g, h ∈ R[M ], then
X X X
((f ∗ g) ∗ h)(m) = (f ∗ g)(x)h(c) = f (a)g(b)h(c)
xc=m xc=m ab=x
X
= f (a)g(b)h(c)
abc=m
X X X
= f (a)g(b)h(c) = f (a)(g ∗ h)(y) = (f ∗ (g ∗ h))(m).
ay=m bc=y ay=m

A special case of this construction which is important in the representation theory
of finite groups is the ring k[G], where k is a field and G is a finite group.

Now suppose that M is an infinite semigroup. Unless we have some sort of ex-
tra structure on R which allows us to deal with convergence of sums – and, in
this level of generality, we do not – the above definition of the convolution product
f ∗ g is problematic because the sum might be infinite. For instance, if M = G
P any group, −1
is then our previous definition of (f ∗ g)(m) would come out to be
x∈G f (x)g(x m), which is, if G is infinite, an infinite sum.

Our task therefore is to modify the construction of the convolution product so


as to give a meaningful answer when the semigroup M is infinite, but in such a way
that agrees with the previous definition for finite M .

Taking our cue from the infinite direct sum, we restrict our domain: define R[M ] to
be subset of all functions f : M → R such that f (m) = 0 except for finitely many
m (or, for short, finitely nonzero functions). Restricting to such functions,
X
(f ∗ g)(m) := f (a)g(b)
ab=m
makes sense: although the sum is apparently infinite, all but finitely terms are zero.
Proposition 5.35. Let R be an associative algebra and M a semigroup. The
structure (R[M ], +, ∗) whose underlying set is the set of all finitely nonzero func-
tions from M to R, and endowed with the binary operations of pointwise additition
and convolution product, is an associative algebra. If R is a ring and M is a monoid
with identity element e, then R[M ] is a ring with multiplicative identity the function
I which takes eM to 1R and every other element of M to 0R .
6. SEMIGROUP ALGEBRAS 127

Exercise 5.32. Prove Proposition 5.35. More precisely, verify that the proof
of Proposition 5.34 goes through unchanged.
L
As a commutative group, R[M ] is naturally isomorphic to the direct sum m∈M R,
i.e., of copies of R indexed by M . One can therefore
P equally well view an element
R[M ] as a formal finite expressions of the form m∈M am m, where am ∈ R and
all but finitely many are 0. Written in this form, there is a natural way to define
the product ! !
X X
am m bm m
m∈M m∈M
of two elements f and g of R[M ]: namely we apply distributivity, use the multi-
plication law in R to multiply the am ’s and the bm ’s, use the operation in M to
multiply the elements of M ,Pand then finally use the addition law in R to rewrite
the expression in the form m cm m. But a moment’s thought shows that cm is
nothing else than (f ∗ g)(m). On the one hand, this makes the convolution product
look very natural. Conversely, it makes clear:

The polynomial ring R[t] is canonically isomorphic to the monoidPring R[N]. In-
deed, the explict isomorphism is given by sending a polynomial n an tn to the
function n 7→ an .

This gives a new proof of the associativity of the product in the polynomial ring
R[t]. We leave it to the reader to decide whether this proof is any easier than direct
verification.. Rather the merit is that this associativity computation has been done
once and for all in a very general context.

The semigroup algebra construction can be used to define several generalizations


of the polynomial ring R[t].
Exercise 5.33. For a ring R, identify the monoid ring R[Z] with the ring
R[t, t−1 ] of Laurent polynomials.
L
First, let T = {ti } be a set. Let F A(T ) := i∈T (N, +) be the direct sum of
a number of copies of (N, +) indexed by T . Let R be a ring, and consider the
monoid ring R[F A(T )]. Let us write the composition law in F A(T ) multiplicatively;
moreover, viewing an arbitrary element IQof F A(T ) as a finitely nonzero function
from T to N, we use the notation tI for t∈T tI(t) . Then an arbitrary element of
Pn
R[F A(T )] is a finite sum of the form k=1 rk tIk , where I1 , . . . , Ik are elements of
F A(t). This representation of the elements should make clear that we can view
R[F A(T )] as a polynomial ring in the indeterminates t ∈ T : we use the alternate
notation R[{ti }].
Theorem 5.36. Let R be a ring and G a group. Then the group ring R[G]
is a domain – i.e., a commutative ring without nonzero zero-divisors – iff R is a
domain and G is commutative and torsionfree.
Proof. Step 1: The ring R[G] is commutative iff both R and G are. Since R
is a subring of R[G], if R[G] is a domain then so is R. If there is g ∈ G and n > 1
such that g n = 1, then (g − 1)(g n−1 + . . . + g + 1) = 0, so R[G] is not a domain.
Step 2: Suppose R is a domain and G is torsionfree commutative. Let K be
128 5. EXAMPLES OF RINGS

the fraction field of R. Then R[G] is a subring of K[G], so it is enough to show


that K[G] is a domain. Here is the key observation: for x, y ∈ K[G]• , there is a
finitely generated subgroup H of G such that x, y ∈ K[H]. Since H is also torsion-
free, we have H ∼ = Zn so K[H] is isomorphic to the ring of Laurent polynomials
−1
K[t1 , t1 , . . . , tn , t−1
n ], which itself lies in function field K(t1 , . . . , tn ) so is a domain.
Somewhere in here we must have used that K[t1 , . . . , tn ] is a domain, so that it
has a field of fractions. This is no problem to establish directly: write K[t1 , . . . , tn ] =
K[t1 , . . . , tn−1 ][tn ] to reduce to the case of a polynomial ring in one variable over a
domain. If deg(f ) = d1 and deg(g) = d2 , then deg(f g) = d1 + d2 .) 

Remark 6. Our convention that a domain is a commutative ring saved us from


considering the following question: if R is a ring without (nonzero) zero-divisors
and G is a group without (nontrivial) elements of finite order, is R[G] a ring without
zero divisors? This question remains wide open even in the case when R is a field,
in which case it is known as the Kaplansky Zero Divisor Conjecture.
Let us go back to the monoid ring R[N], whose elements are finitely nonzero func-
tions f : R → N. Notice that in this case the precaution of restricting finitely
nonzero functions is not necessary: the monoid (N, +), although infinite, has the
property that for any m ∈ N, the set of all x, y ∈ N such that x + y = m is finite
(indeed, of cardinality m + 1). Let us call an arbitrary monoid M divisor-finite
if for each m in M , the set {(x, y) ∈ M 2 | xy = m} is finite.
L
Exercise 5.34. a) For a set T , F A(T ) = t∈T (N, +) is divisor-finite.
b) A group is divisor-finite iff it is finite.
For a divisor-finite monoid M , and any ring R, we may define the big monoid
ring R[[M ]] to be the collection of all functions M → R, with pointwise addition
and convolution product.

For example, if M = (N, +), then writing M multiplicatively with n ∈ N 7→ tn for


some formal generator t, an element of the ring R[[M ]] is an infinite formal sum
n
P
n∈N rn t . Such sums are added coordinatewise and multiplied by distributivity:

X X n
X
( rn tn )( sn tn ) = r0 s0 + (r0 s1 + r1 s0 )t + . . . + ( rk sn−k )tn + . . . .
n∈N n∈N k=0

This ring is denoted by R[[t]] and called the formal power series ring over R.
Exercise 5.35. Using Exercise 5.36, define, for any set T = {ti } and any ring
R, a formal power series ring R[[{ti }]].
Here is yet another variation on the construction: suppose M is a commutative,
cancellative divisor-finite monoid endowed with a total order relation ≤. (Example:
(N, +) or F A(T ) for any T .) There is then a group completion G(M ) together with
an injective homomorphism of monoids M → G(M ). If M is finite and cancellative,
it is already a group. If M is infinite, then so is G(M ), so it cannot be divisor-finite.
Nevertheless, the ordering ≤ extends uniquely to an ordering on G(M ), and we can
define a ring R((G(M )) whose elements are the functions from f : G(M ) → R
such that {x ∈ G(M ) | x < 0, f (x) 6= 0} is finite, i.e., f is finitely nonzero on the
negative values of G(M ).
6. SEMIGROUP ALGEBRAS 129

Exercise 5.36. a) Show: under the above hypotheses, the convolution product
on R((G(M )) is well-defined, and endows R((G(M )) with the structure of a ring.
b) When M = (N, +), identify R((M )) as R((t)), the ring of formal finite-tailed
Laurent series with coefficients in R. Give a multi-variable analogue of this by
taking M = F A(T ) for arbitrary T .
Exercise 5.37. Let R be a not-necessarily-commutative ring. Give a rigorous
definition of the ring Rht1 , t2 i of “noncommutative polynomials” – each ti commutes
with each element of R, but t1 and t2 do not commute – as an example of a small
monoid ring R[M ] for a suitable monoid M . Same question but with an arbitrary
set T = {ti } of noncommuting indeterminates.
The universal property of semigroup rings: Fix a commutative ring R. Let B
be a commutative R-algebra and M a commutative monoid. Let f : R[M ] → B be
an R-algebra homomorphism. Consider f restricted to M ; it is a homomorphism
of monoids M → (B, ·). Thus we have defined a mapping
HomR-alg (R[M ], B) → HomMonoid (M, (B, ·)).
Interestingly, this map has an inverse. If g : M → B is any homomorphism satisfy-
ing g(0) = 0, g(m1 + m2 ) = g(m
P 1 ) + g(m2 ), then
P g extends to a unique R-algebra
homomorphism R[M ] → B: m∈M rm m →
7 m rm g(m). The uniqueness of the
extension is immediate, and that the extended map is indeed an R-algebra homo-
morphism can be checked directly (please do so).

In more categorical language, this canonical bijection shows that the functor M 7→
R[M ] is the left adjoint to the forgetful functor (S, +, ·) 7→ (S, ·) from R-algebras
to commutative monoids. Yet further terminology would express this by saying
that R[M ] is a “free object” of a certain type.
Theorem 5.37. (Universal property of polynomial rings) Let T = {ti } be a set
of indeterminates. Let R be a commutative ring, and S an R-algebra. Then each
map of sets T 7→ S extends to a unique R-algebra homomorphism R[T ] → S.
Proof:
L By the previous result, each monoid map from the free commutative monoid
t∈T Z to S extends to a unique R-algebra homomorhpism. So what is needed is
L map T → M to a commutative monoid extends uniquely
the fact that every set
to a homomorphism t∈T Z → M (in other words, we pass from the category of
sets to the category of commutative R-algebras by passing through the category
of commutative monoids, taking the free commutative monoid associated to a set
and then the free R-algebra associated to the monoid). As before, the uniqueness
of the extension is easy to verify.
Exercise 5.38. a) Formulate analogous universal properties for Laurent poly-
nomial rings, and non-commutative polynomial rings.
b) Suppose M is a divisor-finite monoid. Is there an analogous extension property
for the big monoid ring R[[M ]]?
This result is of basic importance in the study of R-algebras. For instance, let S
be an R-algebra. A generating set for S, as an R-algebra, consists of a subset T of
S such that the least R-subalgebra of S containing T is S itself. This definition is
not very concrete. Fortunately, it is equivalent to the following:
130 5. EXAMPLES OF RINGS

Theorem 5.38. Let R be a commutative ring, S a commutative R-algebra, and


T a subset of S. the following are equivalent:
(i) T generates S as an R-algebra.
(ii) The canonical homomorphism of R-algebras R[T ] → S – i.e., the unique one
sending t 7→ t – is a surjection.
Exercise 5.39. Prove Theorem 5.38.
In particular, a commutative R-algebra S is finitely generated iff it is a quotient
ring of some polynomial ring R[t1 , . . . , tn ].

Another application is that every commutative ring whatsoever is a quotient of


a polynomial ring (possibly in infinitely many indeterminates) over Z. Indeed, for
a ring R, there is an obvious surjective homomorphism from the polynomial ring
Z[R] – here R is being viewed as a set of indeterminates – to R, namely the one
carrying r 7→ r.

A ring R is said to be absolutely finitely generated if it is finitely generated as


a Z-algebra; equivalently, there exists an n ∈ N and an ideal I in Z[t1 , . . . , tn ] such
that Z[t1 , . . . , tn ] is isomorphic to R.
Exercise 5.40. a) Show: a finitely generated ring has finite or countably infi-
nite cardinality.
b) Find all fields which are finitely generated as rings.
(N.B.: In field theory there is a notion of absolute finite generation for a field. This
a much weaker notation: e.g. Q(x) is absolutely finitely generated as a field but not
as a ring.)
CHAPTER 6

Swan’s Theorem

We now digress to discuss an important theorem of R.G. Swan on projective mod-


ules over rings of continuous functions.

Throughout this section K denotes either the field R or the field C, each endowed
with their standard Euclidean topology. For a topological space X, the set C(X)
of all continuous functions f : X → K forms a commutative ring under pointwise
addition and multiplication.

1. Introduction to (topological) vector bundles


1
Recall the notion of a K-vector bundle over a topological space X. This is given
by a topological space E (the “total space”), a surjective continuous map π : E → X
and on each fiber Ex := π −1 (x) the structure of a finite-dimensional K-vector space
satisfying the following local triviality property: for each x ∈ X, there exists an
open neighborhood U containing x and a homeomorphism f : π −1 U → U × K n
such that for all y ∈ U f carries the fiber Ey over y to {y} × K n and induces on
these fibers an isomorphism of K-vector spaces. (Such an isomorphism is called
a local trivialization at x.) As a matter of terminology we often speak of “the
vector bundle E on X” although this omits mention of some of the structure.

On any K-vector bundle E over X we have a rank function r : X → N, namely


we define r(x) to be the dimension of the fiber Ex . We say that E is a rank n
vector bundle if the rank function is constantly equal to n. The existence of local
trivializations implies that the rank function is locally constant – or equivalently,
continuous when N is given the discrete topology, so if the base space X is con-
nected the rank function is constant.

As a basic and important example, for any n ∈ N we have the trivial rank n
vector bundle on X, with total space X × K n and such that π is just projection
onto the first factor.

If π : E → X and π 0 : E 0 → X are two vector bundles over X, a morphism


of vector bundles f : E → E 0 is a continuous map of topological spaces from E to
E 0 over X in the sense that π = π 0 ◦ f – equivalently f sends the fiber Ex to the
fiber Ex0 – and induces a K-linear map on each fiber. In this way we get a category
Vec(X) of K-vector bundles on X. If we restrict only to rank n vector bundles and
morphisms between them we get a subcategory Vecn (X). A vector bundle E on X
is said to be trivial (or, for emphasis, “globally trivial”) if it is isomorphic to the

1from a previous life, if necessary

131
132 6. SWAN’S THEOREM

trivial rank n vector bundle for some n.

Many of the usual linear algebraic operations on vector spaces extend immediately
to vector bundles. Most importantly of all, if E and E 0 are two vector bundles on
X, we can define a direct sum E ⊕E 0 , whose definining property is that its fiber over
each point x ∈ X is isomorphic to Ex ⊕ Ex0 . This not being a topology/geometry
course, we would like to evade the precise construction, but here is the idea: it is
obvious how to define the direct sum of trivial bundles. So in the general case, we
define the direct sum by first restricting to a covering family {Ui }i∈I of simultane-
ous local trivializations of E and E 0 and then glue together these vector bundles
over the Ui ’s. In a similar way one can define the tensor product E ⊗ E 0 and the
dual bundle E ∨ .

For our purposes though the direct sum construction is the most important. It
gives Vec(X) the structure of an additive category: in addition to the existence
of direct sums, this means that each of the sets Hom(E, E 0 ) of morphisms from E
to E 0 form a commutative group. (In fact Hom(E, E 0 ) naturally has the structure
of a K-vector space.) Decategorifying, the set of all isomorphism classes of vector
bundles on X naturally forms a commutative monoid under direct sum (the iden-
tity is the trivial vector bundle X → X where each one point fiber is identified –
uniquely! – with the zero vector space). The Grothendieck group of this monoid is
K(X): this is the beginning of topological K-theory.

2. Swan’s Theorem
But we digress from our digression. A (global) section of a vector bundle π : E →
X is indeed a continuous section σ of the map π, i.e., a continuous map σ : X → E
such that π ◦ σ = 1X . The collection of all sections to E will be denoted Γ(E).
Again this is a commutative group and indeed a K-vector space, since we can add
two sections and scale by elements of K.
But in fact more is true. The global sections form a module over the ring
C(X) of continuous K-valued functions, in a very natural way: given a section
σ : X → E and f : X → K, we simply define f σ : X → E by x 7→ f (x)σ(x). Thus
Γ : E → Γ(E) gives a map from vector bundles over X to C(X)-modules.
In fancier language, Γ gives an additive functor from the category of vector
bundles on X to the category of C(X)-modules; let us call it the global section
functor. (Indeed, if we have a section σ : E → X of E and a morphism of vector
bundles f : E → E 0 , f (σ) = f ◦ σ is a section of E 0 . No big deal!)
Theorem 6.1. (Swan [Sw62]) Let X be a compact space. Then the global
section functor Γ gives an equivalence of categories from Vec(X) to the category of
finitely generated projective C(X)-modules.
In other words, at least for this very topologically influenced class of rings C(X),
we may entirely identify finitely generated projective bundles with a basic and im-
portant class of geometric objects, namely vector bundles.

There is a special case of this result which is almost immediately evident. Namely,
suppose that E is a trivial vector bundle on X, i.e., up to isomorphism E is simply
X × K n with π = π1 . Thus a section σ is nothing else than a continuous function
3. PROOF OF SWAN’S THEOREM 133

σ : X → K n , which in turn is nothing else than an n-tuple (f1 , . . . , fn ) of elements


of C(X). Thus if we define σi ∈ Γ(E) simply to be the section which takes each
point to the ith standard basis vector ei of K n , we see immediately that (σ1 , . . . , σn )
is a basis for Γ(E), i.e., Γ(E) is a free C(X)-module of rank n. Moreover, we have
Hom(X × K n , X × K m ) ∼ = Map(X, HomK (K n , K m ))
= C(X) ⊗K Hom(K n , K w ) ∼
∼ = HomC(X) (Γ(X × K n ), Γ(X × K m )).
Thus we have established that Γ gives an additive equivalence from the category of
trivial vector bundles on X to the category of finitely generated free C(X)-modules.
We wish to promote this to an equivalence from locally trivial vector bundles (i.e.,
all vector bundles) to finitely generated projective modules. Oh, if only we had
some nice “geometric” characterization of finitely generated projective modules!

But we do: namely Proposition 3.11 characterizes finitely generated projective


modules over any commutative ring R as being precisely the images of projection
operators on finitely generated free modules. Thus the essence of what we want
to show is that for any vector bundle E over X (a compact space), there exists a
trivial vector bundle T and a projection P : T → T – i.e., an element of Hom(T, T )
with P 2 = P such that the image of P is a vector bundle isomorphic to E. Indeed,
if we can establish this, then just as in the proof of 3.11 we get an internal direct
sum decomposition T = P (T ) ⊕ (1 − P )(T ) and an isomorphism P (T ) ∼ = E, and
applying the additive functor Γ this gives us that Γ(E) is isomorphic to a direct
summand of the finitely generated free C(X)-module Γ(T ). A little thought shows
that in fact this proves the entire result, because we have characterized Vec(X) as
the “projection category” of the additive category trivial vector bundles, so it must
be equivalent to the “projection category” of the equivalent additive category of
finitely generated free C(X)-modules. So from this point on we can forget about
projective modules and concentrate on proving this purely topological statement
about vector bundles on a compact space.2

3. Proof of Swan’s Theorem


Unfortunately the category of vector bundles over X is not an abelian category. In
particular, it can happen that a morphism of vector bundles does not have either a
kernel or image. Swan gives the following simple example: let X = [0, 1], E = X×K
the trivial bundle, and f : E → E be the map given by f (x, y) = (x, xy). Then
the image of f has rank one at every x 6= 0 but has rank 0 at x = 0. Since X is
connected, a vector bundle over X should have constant rank function. Exactly
the same considerations show that the kernel of f is not a vector bundle. However,
nothing other than this can go wrong, in the following sense:
Proposition 6.2. For a morphism f : E → E 0 of vector bundles over X, the
following are equivalent:
(i) The image of f is a subbundle of E 0 .
(ii) The kernel of f is a subbundle of E.
2We note that [Sw62] takes a more direct approach, for instance proving by hand that
the global section functor Γ is fully faithful. In our use of projection operators and projection
categories to prove Swan’s theorem we follow Atiyah [At89, §1.4]. Aside from being a bit shorter
and slicker, this approach really brings life to Proposition 3.11 and thus seems thematic in a
commutative algebra course. But it is not really more than a repackaging of Swan’s proof.
134 6. SWAN’S THEOREM

(iii) The function x 7→ dimK (Im f )x is locally constant.


(iv) The function x 7→ dimK (Ker f )x is locally constant.
Proof. Step 1: We first wish to prove a special case: namely that if f : E → E 0
is a monomorphism of vector bundles (i.e., it induces an injection on all fibers) then
(Im f ) is a subbundle of E 0 and f : E → (Im f ) is an isomorphism. The issues of
whether Im f is a vector bundle and f is an isomorphism are both local ones, so it
suffices to treat the case where E and E 0 are trivial bundles. Suppose E 0 = X × V ,
and let x ∈ X. Choose Wx ⊂ V a subspace complementary to (Im f )x . Then
G := X × Wx is a sub-bundle of E; let ι : G → E be the inclusion map. Define
θ : E ⊕ G → E 0 by θ((a, b)) = f (a) + ι(b). Then θx is an isomorphism, so there
exists an open neighborhood U of x such that θ|U is an isomorphism. Since E is a
subbundle of E ⊕ G, θ(E) = f (E) is a subbundle of θ(E ⊕ G) = E 0 on U .
Step 2: Since the rank function on a vector bundle is locally constant, (i) =⇒
(iii), (ii) =⇒ (iv), and (by simple linear algebra!) (iii) ⇐⇒ (iv).
(iv) =⇒ (i): Again the issue of whether Im f is a vector bundle is a local one,
so we may assume that E = X × V is a trivial bundle. For x ∈ X, let Wx ⊂ V
be a complementary subspace to (Ker f )x . Let G = X × Wx , so that f induces
a homomorphism ψ : G → E 0 whose fiber at x is a monomorphism. Thus ψ is a
monomorphism on some neighborhood U of x, so ψ(G)|U is a subbundle of E 0 |U .
However Ψ(G) ⊂ f (E), and since f (E) has constant rank, and
dim ψ(G)y = dim ψ(G)x = dim f (E)x = dim f (E)y
for all y ∈ U , ψ(G)|U = f (E)|U . so f (E) is a subbundle of E 0 .
(iv) =⇒ (ii): here we exploit dual bundles. The hypothesis implies that the
kernel of f ∨ : (E 0 )∨ → E ∨ has constant rank function. Since E ∨ → Coker f ∨ is
an epimorphism, (Coker f ∨ )∨ → E ∨∨ is a monomorphism: by Step 1, its image is
a subbundle. But the natural map E → E ∨∨ is an isomorphism, the restriction of
which to Ker f gives an isomorphism to the vector bundle (Coker f ∨ )∨ . So Ker f
is a vector bundle. 
The proof yields the following additional information.
Corollary 6.3. For any morphism of vector bundles, the rank function of the
image is upper semi-continuous: that is, for any x ∈ X, there exists a neighborhood
U of x such that for all y ∈ U , dimK (Im f )y ≥ dimK (Im f )x .
Exercise 6.1. Prove Corollary 6.3.
Proposition 6.4. Let E be a vector bundle over X, and let P ∈ End(E) =
Hom(E, E) be a projection, i.e., P 2 = P . Then:
a) We have Ker(P ) = Im(1 − P ).
b) Im P and Im(1 − P ) are both subbundles of E.
c) There is an internal direct sum decomposition E = Im P ⊕ Im(1 − P ).
Proof. a) For all x ∈ X linear algebra gives us an equality of fibers Ker(P )x =
Im(1 − P )x . This suffices!
b) From part a) we deduce an equality of rank functions
rIm P + rIm(1−P ) = rE .
By Corollary 6.3, for all x ∈ X, there is a neighborhood U of x on which rIm P is
at least as large as rIm P (x), rIm(1−P ) is at least as large as rIm(1−P ) (x) and rE is
3. PROOF OF SWAN’S THEOREM 135

constantly equal to rE (x). On this neighborhood the ranks of Im P and Im(1 − P )


must be constant, and therefore by Proposition 6.2 Im P and Im(1 − P ) are both
subbundles.
c) Again it is enough to check this fiber by fiber, which is simple linear algebra. 
An inner product on a finite-dimensional R-vector space V is, as usual, a symmetric
R-bilinear form h, i : V × V → R which is positive definite in the sense that for
all x ∈ V \ {0}, hx, xi = 0. An inner product on a finite-dimensional C-vector
space V is a positive definite sesquilinear form: i.e., it is C-linear in the first vari-
able, conjugate-linear in the second variable and again we have hx, xi > 0 for all
x ∈ V \ {0}.

Now let E be a K-vector bundle on X. An inner product on E is a collection of


inner products h, ix : Ex × Ex → K on each of the fibers which vary continuously in
x. Formally this means the following: let E ×X E be the subset of (e1 , e2 ) ∈ E × E
such that π(e1 ) = π(e2 ); then such a fiberwise family of inner products defines a
function from E ×X E to K, and this function is required to be continuous.

Let us say that a metrized vector bundle E on X is a vector bundle together


with an inner product. (Again, this is an abuse of terminology: we do not speak of
the inner product by name.)
Proposition 6.5. Let E be a metrized line bundle on X.
a) If E 0 is a subbundle of E, fiberwise orthogonal projection onto E 0 defines a
projection operator P ∈ End(E) with image E 0 .
b) All short exact sequences 0 → E 0 → E → E 00 → 0 of vector bundles are split.
c) If M is another vector bundle on X and there exists an epimorphism of bundles
q : E → M , then M is isomorphic to the image of a projection operator on E.
Proof. a) This is mostly a matter of understanding and unwinding the defi-
nitions, and we leave it to the reader.
b) Let P be orthogonal projection onto E 0 . The restriction of the map E → E 00 to
Ker P is an isomorphism of vector bundles. The inverse of this isomorphism gives
a splitting of the sequence.
c) By Proposition 6.2, since Im q = M is a vector bundle, so is Ker q, whence a
short exact sequence
0 → Ker q → E → M → 0.
By part b), there exists a splitting σ : M → E of this sequence. Then, as usual,

P = σ ◦ q is a projection operator on E and q|Im P : Im P → M . 
Proposition 6.6. If X is a paracompact topological space, then every vector
bundle over X admits an inner product.
Proof. This is a rather standard topological argument which we just sketch
here. Let M be a vector bundle on X, and let {Ui }i∈I be an open covering of X
such that the restriction of M to each Ui is a trivial bundle. On a trivial bundle
there is an obvious inner product, say h, ix . Now, since X is paracompact, there
exists a partition of unity {ϕi }i∈I subordinate to the open covering {Ux }: that is,
• each ϕi : X → [0, 1] is continuous,
• for all x ∈ X we have supp(ϕi ) ⊂ Ui
• for all x ∈ X there exists an open neighborhood V of x on which all but finitely
136 6. SWAN’S THEOREM

many ϕi ’s vanishPidentically, and


• for all x ∈ X, i∈I ϕi (x) = 1.3
Then, for x ∈ X and e1 , e2 ∈ Mx , define
X
he1 , e2 ix := ϕi (x)he1 , e2 ii ;
i
the sum extends over all i ∈ I such that x ∈ Ui . This is an inner product on M . 
To complete the proof of Swan’s Theorem, it suffices to show that if X is com-
pact, every vector bundle M on X is the epimorphic image of a trivial bundle. In
particular, Proposition X.X then shows that M is a direct summand of a trivial
vector bundle T and thus Γ(M ) is a direct summand of the finitely generated free
C(X)-module Γ(T ), hence is finitely generated projective.
Proposition 6.7. Let X be a compact space and M a vector bundle on X.
Then there exists an epimorphism of bundles from a trivial vector bundle X × V to
M.
Proof. Step 1: We claim that for each x ∈ X, there exists a neighborhood
Ux of x and finite set of global sections Sx = {sx,1 , . . . , sx,kx } of M such that for
all y ∈ U , sx,1 (y), . . . , sx,kx (y) is a K-basis for My .
proof of claim: Let U be an open neighborhood of x on which M is a trivial
bundle. Certainly then there exist finitely many sections s1 , . . . , sn of M over U
which when evaluated at any y ∈ U give a basis of My . We need to show that there
exists an open set W with x ∈ W ⊂ U and global sections s01 , . . . , s0n such that
for all i, s0i |W = si |W . For this it suffices to work one section at a time: let s be
a section of M over U . Since X is paracompact, it is normal, so there exist open
neighborhoods W and V of x with W ⊂ V , V ⊂ U . By Urysohn’s Lemma, there is
a continuous function ω : X → [0, 1] such that ω|W ≡ 1 and ω|X\V ≡ 0. If we then
define s0 : X → M by s0 (y) = ω(y)s(y) for y ∈ U and s0 (y) = 0 for y ∈ X \ U , then
this s does the job.
Step 2: By compactness S of X, there exists a finite subset I of X such that {Ux }x∈I
covers X. So S = i∈I Sx is a finite set of global sections of M which when
evaluated at any x ∈ X, span the fiber Mx . So the K-subspace V of Γ(M ) spanned
by S is finite-dimensional. We define a map q : X × V → M by q(x, s) = s(x).
This is a surjective bundle map from a trivial vector bundle to M ! 
Remark: In the above proof the paracompactness of X seems to have been fully
exploited, but the need for compactness is less clear. In fact, at the end of [Sw62],
Swan remarks that if you replace the last step of the proof by an argument from
Milnor’s 1958 lecture notes Differential Topology, one gets a categorical equivalence
between vector bundles with bounded rank function on a paracompact space X and
finitely generated projective C(X)-modules.

Remark: A more straightforward variant of Swan’s theorem concerns the case where
X is a compact differentiable manifold (say of class C ∞ ). In this case the equiva-
lence is between differentiable K-vector bundles on X and modules over the ring of
K-valued C ∞ -functions. Looking over the proof, one sees that the only part that
needs additional attention is the existence of differentiable partitions of unity. Such
3See e.g. Exercise 5 in §4.5 of Munkres’ Topology: a first course for a proof of this fact.
5. STABLY FREE MODULES 137

things indeed exist and are constructed in many of the standard texts on geometry
and analysis on manifolds. We recommend [We80], which has a particularly clear
and complete discussion.

4. Applications of Swan’s Theorem


4.1. Vector bundles and homotopy.

Vector bundles on a space are of interest not only to differential topologists and
geometers but also to algebraic geometers. This is because pullback of vector bun-
dles behaves well under homotopy.

First, suppose that f : X → Y is a continuous map of topological spaces and


π : E → Y is a vector bundle on Y . We may pullback π to a vector bundle
πX : E ×Y X → X just by taking E ×Y X to be the fiber product of the maps
f and π, namely the subspace of X × E consisting of all pairs (x, v) such that
f (x) = π(v) ∈ Y . The map πX : E ×Y X → X is just restriction of the projection
map: (x, v) 7→ x.
Exercise 6.2. Show: πX : E ×Y X → X is indeed a vector bundle on X. We
abbreviate it by either f ∗ π or (more abusively) f ∗ E.
Exercise 6.3. Show: the pullback of a trivial bundle is a trivial bundle.
Theorem 6.8. (Covering Homotopy Theorem) Let X and Y be topological
spaces with X paracompact. Let π : E → Y be a vector bundle on Y , and let
f, g : X → Y be homotopic maps. Then the pullbacks f ∗ π and g ∗ π are isomorphic
vector bundles on X.
Proof. See for instance [Hs66, Thm. 4.7]. 
For our applications, it is enough to know that compact spaces are paracompact.
But for culture we also remark that any regular σ-compact space is paracompact,
e.g. any CW-complex with only finitely many cells of any given dimension.
Corollary 6.9. If X is a contractible paracompact space, then every vector
bundle on X is trivial.
Proof. Choose any point x0 ∈ X, let f : X → X be the map which sends
every point of X to x0 ,and let g : X → X be the identity map. If π : E → X is
any vector bundle on X, then by Theorem 6.8 we have f ∗ π = g ∗ π. Since g is the
identity map, g ∗ π = π. On the other hand, tracking through the definitions shows
f ∗ π = X × π −1 (x0 ), a trivial bundle. So π is trivial. 

5. Stably free modules


Recall that an R-module M is stably free if there is a finitely generated free
module F such that M ⊕ F is free. This definition is natural from the perspective
^
of K-theory: the class [P ] in K0 (R) of a finitely generated projective module P is
trivial iff P is stably free.
Exercise 6.4. Let 0 → A → B → P → 0 be a short exact sequence of R-
modules, with P stably free. Show: A is stably free ⇐⇒ B is stably free.
138 6. SWAN’S THEOREM

Certainly we have
free =⇒ stably free =⇒ projective .
Asking to what extent these implications can be reversed brings us quickly to some
deep and beautiful mathematics.
5.1. Finite Generation.

We begin by addressing finite generation conditions.


Exercise 6.5. (Eilenberg Swindle): Let us say that a projective module P is
weakly stably free if there exists a not necessarily finitely generated free module
F such that P ⊕ F is free. Show that every projective module is weakly stably free.
(Hint: if P ⊕ Q is free, take F = P ⊕ Q ⊕ P ⊕ Q ⊕ . . ..)
Exercise 6.6. Show: for a finitely generated projective module P , the following
are equivalent:
(i) P is stably free.
(ii) P admits a finite free resolution: for some n ∈ N there is an exact sequence
0 → Fn → . . . → F0 → P → 0,
with each Fi a finitely generated free module.
This explains why the free module we take the direct sum with in the definition of
stably free is required to be finitely generated. What happens if we take the module
P to be infinitely generated? Here let us be sure that by an infinitely generated
R-module, we mean an R-module which is not finitely generated.4
Theorem 6.10. (Gabel) Each infinitely generated stably free module is free.
Proof. Let M be infinitely generated and stably free. Choose a ∈ N such that
F = M ⊕Ra is free. Let {ai }i∈I be a basis for F . Since F has an infinitely generated
homomorphic image, I is infinite. Let p : F → Ra be the natural projection map
(x, y) 7→ y. For each standard basis element ek of Ra lift it to eek in F and let Jk
be the “support” of eekS
, i.e., the set of indices i such that the coefficient of ai in eek
a
is nonzero. Then J = k=1 Jk is finite. Let F 0 = hai ii∈J , so that F 0 is free of finite
rank and F/F 0 is free of infinite rank. By construction q(F 0 ) = Ra ; it follows that
F = F 0 + M.
Put N = M ∩ F 0 , so
F 0 /N ∼
= Ra .
Since Ra is projective, the sequence
0 → N → F 0 → Ra → 0
splits, giving
F0 ∼
= N ⊕ Ra .
Further
F/F 0 ∼
= M/N,
4A priori it would be reasonable to take “infinitely generated R-module” to mean a module
which possesses an infinite generating set, but a moment’s thought shows that an R-module has
this property iff it is infinite, so it is more useful to define “infinitely generated” as we have.
5. STABLY FREE MODULES 139

so M/N is infinitely generated free: M/N ∼ = Ra ⊕ F 00 for a free module F 00 . In


particular M/N is projective, so the sequence
0 → N → M → M/N → 0
splits. Putting all this together we get
M∼ = N ⊕ M/N ∼ = N ⊕ Ra ⊕ F 00 ∼
= F 0 ⊕ F 00 . 
The following result – which we will not prove here – shows that for a large class of
“reasonable rings” infinitely generated projective modules are much less interesting
objects than finitely generated projectives, and thus gives further motivation to our
restriction to the finitely generated case.
Theorem 6.11. (Bass [Ba63]) Let R be a connected Noetherian ring. Then
any infinitely generated projective R-module is free.
However, we can use Swan’s Theorem to exhibit a nonfree infinitely generated pro-
jective module. Let [0, 1] be the closed unit interval with its topology: a compact,
contractible space. By Corollary 6.9, every vector bundle over [0, 1] is trivial. By
Swan’s Theorem, this implies that every finitely generated projective module over
the ring R = C([0, 1]) of continuous functions f : [0, 1] → R is free.

But now – as in §3.9 – consider the ideal I of all functions f ∈ R which van-
ish near zero, i.e., for which there exists (f ) > 0 such that f |[0,(f )] ≡ 0. By
Exercise X.X, I is a projective R-module. Moreover, I is not a free R-module:
indeed, any f ∈ I is annihilated by any continuous function with support lying in
[0, (f )], and nonzero such functions clearly exist. On the other hand, any nonzero
free module has elements with zero annihilator: take any basis element.

Thus C([0, 1]) is a connected ring over which every finitely generated projective
module is free, but the infinitely generated projective module I is not free. (Theo-
rem 6.11 says that no such modules exist over connected Noetherian rings.) More-
over I is therefore clearly not a direct sum of finitely generated modules, since by
what we have established any such module over C([0, 1]) would be free!
Exercise 6.7. Use Corollary 3.49 to give a purely algebraic proof that I is not
a direct sum of finitely generated submodules.
Exercise 6.8. Find necessary and sufficient conditions on a compact, con-
tractible space X for there to exist a nonfree projective module.
5.2. Ranks.

Later we will attach to a finitely generated projective module over any ring R
a rank function (on Spec R). However, for a stably free module we can – as for free
modules – simply assign a rank. Namely, if we put rank P = b − a.
Exercise 6.9. Show: the rank of a finitely generated stably free module is
well-defined.
Exercise 6.10. Show: for an R-module M , the following are equivalent:
(i) M is stably free of rank zero.
(ii) M = 0.
Comment: This will be quite routine once we have the theory of localization. If you
140 6. SWAN’S THEOREM

have trouble with the general case now, just show that M ⊕ R ∼= R =⇒ M = 0,
which is easier: every cyclic module is isomorphic to R/I for some ideal I of R;
now consider annihilators.
5.3. The least number of generators.

For a finitely generated R-module M we denote by mg M the minimal number


of generators of M , i.e., the least n such that Rn  M .
From a naive perspective this is perhaps the most natural numerical invariant
associated to a finitely generated R-module. But in fact it behaves badly. Essen-
tially its only good property is the obvious one: if M1  M2 , then mg(M2 ) ≤
mg(M1 ). However, if M1 ,→ M2 , then we certainly need not have mg(M1 ) ≤
mg(M2 ): let I be a finitely generated but nonprincipal ideal, and let M1 = I,
M2 = R. One may momentarily hope that for finitely generated R-modules M1
and M2 we at least have mg(M1 ⊕ M2 ) = mg(M1 ) + mg(M2 ) but in fact this is
false even over the simplest rings: take R = Z, M1 = Z/2Z, M2 = Z/3Z. But it
gets even worse:
Exercise 6.11. Let R be a ring and M1 , M2 be finitely generated R-modules.
a) Suppose R is local. Show: mg(M1 ⊕ M2 ) = mg(M1 ) + mg(M2 ). In fact, show
that if 0 → M 0 → M → M 00 → 0 is a short exact sequence of finitely generated
R-modules then mg(M ) = mg(M 0 ) + mg(M 00 ).
b) Suppose R is a PID. Show: mg(M1 ⊕ M1 ) = 2 mg(M1 ).
c) Suppose R is a Dedekind domain,5 and l I is a nonzero proper ideal of R. Show:
(i) • If I is principal, mg(I) = 1.
• If I is not principal, mg(I) = 2.
(ii) • If I 2 is principal, mg(I ⊕ I) = 2.
(iii) • If I 2 is not principal, mg(I ⊕ I) = 3.
d) Deduce: For a nonprincipal ideal I in a Dedekind domain R, mg(I ⊕ I) <
2 mg(I).
Later we will see “better” invariants for certain subclasses of finitely generated R-
modules, namely the rank for projective modules and the length for...finite length
modules. Over a Dedekind domain every finitely generated module can be decom-
posed into the direct sum of a projective module and a finite length module. This
does not hold over more general rings, e.g. the C[x, y]-module C[x] is a torsion
module of infinite length so cannot be so expressed.
Proposition 6.12. A rank one stably free module is free.
We will come back to prove this later once we have developed localization.
5.4. Around Hermite’s Lemma.

In number theory and related branches of mathematics one studies sublattices Λ of


the standard integral lattice Zn , i.e., rank n Z-submodules of Zn . Their structure
is surprisingly rich – for instance, the function Ln (k) which counts the number of
index k sublattices of Zn is arithmetically interesting and nontrivial. In particular,
one question that comes up in the study of integer lattices is: which vectors v ∈ Zn
5This exercise is stated now for continuity purposes, but to solve it you will probably want
to use the theory of finitely generated modules over a Dedekind domain detailed in § 20.6.
5. STABLY FREE MODULES 141

can be part of a Z-basis of Zn ? Unlike the answer for modules over a field (all
nonzero vectors), there is an obvious obstruction: for instance there is no basis
(v1 , v2 ) of Z2 with v1 = (2, 0). For if so, the linear transformation T : Z2 → Z2
given by T ((1, 0)) = (2, 0), T ((0, 1)) = v2 = (a, b) has determinant 2b. Since this
is not a unit in Z, T is not invertible, which is a contradiction (make sure you see
why, e.g. by using the universal property of free modules).

This observation can be vastly generalized, as follows: for a domain R and n ∈ Z+ ,


we say v = (x1 , . . . , xn ) ∈ Rn is a primitive vector if v 6= 0 and hx1 , . . . , xn i = R.
Exercise 6.12. Let K be the fraction field of R. Show that v ∈ (Rn )• is
primitive iff Kv ∩ Rn = Rv.
Exercise 6.13. Let R be a domain, and let (b1 , . . . , bn ) be a basis for Rn . Show
that each bi is a primitive vector.
In 1850 Hermite proved that for integer lattices this is the only obstruction.
Proposition 6.13. (Classical Hermite Lemma) For a vector v ∈ Zn , the fol-
lowing are equivalent:
(i) There is M ∈ GLn (Z) with M (e1 ) = v, i.e., the first column of M is v.
(ii) There is a basis for Zn containing v.
(iii) v is a primitive vector.
For a proof of Proposition 6.13 in the classical style, see [GoN, § 1.3.3]. In fact the
methods of module theory allow for a slicker proof of a more general result.
Proposition 6.14.
Let R be a PID. For a vector v ∈ Rn , the following are equivalent:
(i) There is M ∈ GLn (R) with M (e1 ) = v, i.e., the first column of M is v.
(ii) There is a basis for Rn containing v.
(iii) v is a primitive vector.
Proof. Any two bases of Rn are equivalent under GLn (R). So (i) ⇐⇒ (ii).
(ii) =⇒ (iii): If v, v2 , . . . , vn is a basis for Rn and v were not primitive, then we
would have v = αw for some α ∈ R• \ R× . Then w = α1 v expresses w as a K-linear
combination of the basis vectors with a nonintegral coefficient: contradiction.
(iii) =⇒ (ii): Step 1: For a domain R and v ∈ (Rn )• , we claim that v is a primitive
vector iff Rn /hvi is torsionfree.
Proof: Suppose v is not primitive: v = αv 0 for some α ∈ R• \ R× . Then v 0 is
a torsion element of Rn /hvi. Conversely, suppose v is primitive. Ifn = 1 then
hvi = R and the result holds trivially, so assume n ≥ 2. Suppose there is w ∈ Rn
β
and α ∈ R• such that αw = βv for some β ∈ R. Thus w = α v. Since v is primitive,
n
α | β and the image of w in R /hvi is zero.
Step 2: Consider the short exact sequence
0 → hvi → Rn → M → 0,
with M = Rn /hvi. By Step 1, M is a finitely generated torsionfree module over a
PID, so it is free: indeed, tensoring to K and applying linear algebra we see that
M ∼ = Rn−1 . Thus the sequence splits: Rn = hvi ⊕ M 0 , with M 0 ∼ = Rn−1 . Thus if
0 n
v2 , . . . , vn is an R-basis for M , v, v2 , . . . , vn is an R-basis for R . 
142 6. SWAN’S THEOREM

Proposition 6.15. Let R be a commutative ring, and let n ∈ Z+ . the following


are equivalent:
(i) If for an R-module M we have M ⊕ R ∼ = Rn then M is free.
(ii) Every primitive vector v ∈ R is part of a basis for Rn .
n

Proof. We follow a treatment of K. Conrad [Cd-SF]. First we observe that


when n = 1 both (i) and (ii) hold for all R-modules M : indeed, by Exercise X.X,
if M ⊕ R ∼ = R then M = 0, whereas (ii) is completely vacuous in this case.
(i) =⇒P(ii): Assume (i). For v = (v1 , . . . , vn ), w = (w1 , . . . , wn ) ∈ Rn , let
n
v · w = i=1 vi wi . Let a = (a1 , . . . , an ) ∈ Rn be a primitive vector. Observe that
this is equivalent to the existence of b = (b1 , . . . , bn ) ∈ Rn with a · b = 1 and fix
such a b. Consider the R-linear functional f : Rn → R given by v 7→ v · b. Since
f (a) = 1 it is nonzero and thus there is a short exact sequence
f
0 → Ker f → Rn → R → 0.

Since R is projective, this sequence splits, giving Rn ∼


= Ker f ⊕ R. More concretely
a splitting is given by a section σ : R → Rn of f which is determined by mapping
1 ∈ R to any v ∈ Rn with f (v) = 1. Thus 1 7→ a gives an internal direct sum
decomposition
Rn = Ker f ⊕ hai ∼ = Ker f ⊕ R.
By our hypothesis (i), Ker f is free, and if b2 , . . . , bn is a basis for Ker f then
a, b2 , . . . , bn is a basis for Rn containing a.

(ii) =⇒ (i): Let g : M ⊕ R → Rn be an R-module isomorphism. Put a =
(a1 , . . . , an ) = g(0, 1). We claim that a is a primitive vector. If not, there is a
maximal ideal m such that ha1 , . . . , an i ⊂ m. But

g|m(M ⊕Rn ) : mM ⊕ m → (mR)n ,

and g(0, 1) = a ∈ (mR)n , so (0, 1) ∈ mM ⊕ (mR)n , a contradiction. Thus by (ii)


there is a basis a, b2 , . . . , bn of Rn , so that g −1 (a), g −1 (b2 ), . . . , g −1 (bn ) is a basis
of M ⊕ R. For 2 ≤ i ≤ n we write g −1 (bi ) = (xi , ci ). Subtracting off from each
of these vectors a suitable scalar multiple of g −1 (a) = (0, 1) we get a new basis
(0, 1), (x2 , 0), . . . , (xn , 0) of M ⊕ R. Then x2 , . . . , xn is a basis for M . 

Theorem 6.16. For a commutative ring R, the following are equivalent:


(i) For all R-modules M , if M ⊕ R is free, then M is free.
(ii) For all n ∈ Z+ , every primitive vector v ∈ Rn is part of a basis of Rn .
(iii) Every stably free R-module M is free.
A ring satisfying these equivalent conditions will be called an L-Hermite ring.

Proof. Since an infinitely generated stably free module is free (Theorem 6.10),
in parts (i) and (iii) we may – and shall – assume M is finitely generated. Then:
(i) ⇐⇒ (ii) is immediate from Proposition 6.15.
(i) =⇒ (iii): It suffices to show that for all finitely generated modules M and all
a ∈ N, if M ⊕ Ra is free then M is free. We go by induction on n, the case n = 0
being trivial. Suppose the result holds for a ∈ N. Then M ⊕ Ra+1 ∼ = (M ⊕ Ra ) ⊕ R
a
is free, so by (i) M ⊕ R is free, and then by induction M is free.
(iii) =⇒ (i) is immediate. 
5. STABLY FREE MODULES 143

5.5. Swan’s Construction.

For n ∈ N, let
Rn = R[t0 , . . . , tn ]/ht20 + . . . + t2n − 1i.
Exercise 6.14. Show: Rn is a domain iff n ≥ 1.
Consider the map H : Rnn+1 → Rn obtained by taking the dot product of v =
(v1 , . . . , vn+1 ) with t = (t0 , . . . , tn ). For 0 ≤ i ≤ n, let ei be the ith standard basis
vector of Rnn+1 ; then H(ei ) = ti , so the image of H contains ht0 , . . . , tn i = Rn : H
is surjective. Let Pn = Ker H, so we have a short exact sequence
H
0 → Pn → Rnn+1 → Rn → 0.
As above, this sequence splits and since t · t = 1, a canonical section is given by
mapping 1 ∈ Rn to t. In particular
Pn ⊕ R n ∼
= Rn+1 n
so Pn is stably free. When is it free?
Theorem 6.17. (Swan) The stably free Rn -module Pn is free iff n = 0, 1, 3 or
7.
Proof. Step 0: Since P0 = 0, it is free. Moreover P1 has rank 1 so is free
by general principles (Proposition 6.12). But this is overkill: in fact one sees that
−t1 e0 + t0 e1 is a basis for P1 . Similarly one can simply write down bases for P3
and P7 in a concrete manner: we leave this as an exercise for the reader.
Step 1: Suppose n ∈ / {0, 1, 3, 7}. We wish to show that Pn is not free. The key
observation is that it is enough to do so after any base change: that is, if Rn → S
is a ring map and M is an R-module such that M ⊗Rn S is not a free S-module,
then M is not a free R-module. What is the natural base change to make?
Notice that Rn is nothing else than the ring of polynomial functions on the
unit sphere S n ⊂ Rn+1 . For those unitiated in the above jargon we spell it out
more explicitly: every f ∈ R[t0 , . . . , tn+1 ] induces a function from Rn+1 → R
and thus by restriction a function S n → R. We wish to identify polynomials
which define the same function on S n , and to do so we should at least impose the
relation t20 + . . . + t2n − 1 = 0 since this function vanishes identically on S n . As we
will see later when we study the Nullstellensatz, since by Exericse X.X the ideal
I = ht20 + . . . + t2n − 1i is prime, it is radical and thus the relation I(V (I)) = I tells
us that the only polynomials which vanish identically on S n are those in I.
Since every polynomial function is a continuous function for the Euclidean
topology on S n , we get an extension of rings Rn → C(S n ). So our bright idea is to
show insteaad that the finitely generated projective C(S n )-module
Tn = Pn ⊗Rn C(S n )
is not free. By Swan’s Theorem, Tn corresponds to a vector bundle on S n and it is
equivalent to show that this vector bundle is nontrivial. 6
Step 3: We claim that in fact Tn is nothing else but the tangent bundle of S n .
Indeed, we have S n ⊂ Rn+1 . The tangent bundle to Rn+1 is trivial, hence so is its
6Thus in summary we have just accomplished the following exciting maneuver: using ba-
sic affine algebraic geometry, we have completely transferred our problem from the domain of
commutative algebra to that of differential topology!
144 6. SWAN’S THEOREM

pullback to S n , say F n+1 . Further, there is a surjective bundle map from F n+1 to
the rank one trivial bundle F 1 : at every point of S n we orthogonally project to the
outward normal vector. The kernel of this bundle map is T (S n ). Thus we have a
short exact sequence of vector bundles
0 → T S n → F n+1 → F → 0.
We claim that under the Swan’s Theorem equivalence of categories, this split exact
sequence corresponds to the split exact sequence
H
0 → Tn → C(S n )n+1 →→ C(S n ) → 0
which is the base change to C(S n ) of the defining short exact sequence of S n . We
leave it to the interested reader to piece this together from our consruction of Pn .
Step 4: By a classical theorem of Bott and Milnor [BM58], the tangent bundle of
S n is trivial iff n ∈ {0, 1, 3, 7}. 
Exercise 6.15. Show: Pn is free for n = 3 and n = 7.
Exercise 6.16. Find stably free but not free modules of ranks 3 and 7.
The Bott-Milnor Theorem is a deep and celebrated result. Their original proof used
the recently developed tools of midcentury differential topology: Stiefel-Whitney
and Pontrjagin classes, cohomology operations, and so forth. In 1962 J.F. Adams
determined for each n the largest rank of a trivial subbundle of T (S n ) [Ad62]. The
K-theory developed in the 1960’s gave more graceful proofs: we recommend that
the interested reader consult, for instance, [Ka, § V.2].

If one merely wants some values of n for which Pn is not free, one can use much
lower technology. for instance, the Poincaré-Hopf Theorem [Mi, p. 35] implies
that a closed n-manifold which admits a nowhere vanishing vector field (equiva-
lently a trivial rank one subbundle of its tangent bundle; this is much weaker than
the tangent bundle being trivial) must have zero Euler characteristic. The Euler
characteristic of S n is 1 + (−1)n , so it is nonzero for all even n.
CHAPTER 7

Localization

1. Definition and first properties


As we have seen, one way to “simplify” the study of ideals in a ring R is to pass to
a quotient ring R/I: as we have seen, this has the (often useful) effect of “cutting
off the bottom” of the ideal lattice by keeping only ideals J ⊃ I. There is another
procedure, localization, which effects the opposite kind of simplification: given a
prime ideal P of R, there is a ring RP together with a canonical map ι : R → RP
such that ι∗ : I(RP ) → I(R) is an injection whose image is precisely the ideals
J ⊂ P . As usual, ι∗ carries prime ideals to prime ideals. In particular, assuming
only that P is prime, we get a corresponding ideal – rather inelegantly but stan-
dardly denoted P RP – which is the unique maximal ideal of RP . If we can take
P = (0) – i.e., if R is a domain – this means that P RP is the only ideal of RP ,
which is therefore a field. In fact it is nothing else than the quotient field of the
domain R, and – with one exception – all the secrets of localization are already
present in this very familiar special case.

In fact the localization construction is a bit more general than this: given an arbi-
trary ring R (of course commutative with unity!) and an arbitrary multiplicative
subset S of R – this just means that 1 ∈ S and SS ⊂ S – we will define a new ring
RS together with a canonical homomorphism ι : R → RS (for which ι∗ will still be
an injection with explicitly given image). In fact, just as in the case of quotients,
ι satisfies a certain universal mapping property, but let us sacrifice some elegance
for intelligibility by working our way up to this crisp definition.

Indeed, first consider the special case in which R is a domain, with fraction field
F . Then RS will be an extension ring of R, still with fraction field F , which is
obtained by adjoining to R all elements 1s for s ∈ S.

Example: Suppose R = Z, S = {2n }n∈N . Then RS = Z[ 21 ]. Indeed we see that for


any nonzero element f , we can take S to be the multiplicative set consisting of the
powers of f , and then the localization is just R[ f1 ].

What if in the example above, instead of taking the multiplicative subset gen-
erated by 2, we took the multiplicative subset generated by 22 , or 2127 ? Clearly it
k−1
wouldn’t matter: if we have 21k for any k in our subring of Q, we also have 2 2k = 12 .
To generalize this idea, define the saturation S of a multiplicatively closed subset
S of a domain R to be the set {a ∈ R | ∃b ∈ R | ab ∈ S}, i.e., the set of all divisors
of elements of S. The same observation as above shows that RS = RS , so if we like
we can restrict to consideration of saturated multiplicatively closed subsets.

145
146 7. LOCALIZATION

Example, continued: The saturated, multiplicatively closed subsets of Z corre-


spond to (arbitary) subsets P of the prime numbers (exercise!). In particular Z
itself corresponds to P = ∅, Z[ p1 ] corresponds to P = {p}, Q corresponds to the set
of all primes. Most interestingly, fix any prime p and let P be the set of all primes
except p: then the corresponding ring, which is confusingly denoted Z(p) is the
set of all rational numbers of the form xy where p does not divide y. Notice that
such rings are the maximal subrings of Q which are not fields. Moreover, the units
of Z(p) are precisely the elements of the form xy with (p, x) = 1. The nonunits a
are all of the form pa0 for a0 ∈ Z(p) , so therefore the unique maximal ideal is the
principal ideal (p) = pZ(p) .
Exercise 7.1. Show: the only ideals in Z(p) are those of the form (p)k for
some k ∈ N.
Now let R be any ring and S a multiplicatively closed subset of R. We would still
like to define a ring S −1 R which is, roughly speaking, obtained by adjoining to R
all inverses of elements of S. We can still define S −1 R in terms of formal quotients,
i.e., as equivalence classes of elements (a, b) with a ∈ R, b ∈ S. However, if we
define (a, b) ∼ (c, d) to be ad = bc, then unfortunately we find that this need not
be an equivalence relation! Therefore we need to enlarge the relation a bit: we put
(a, b) ∼ (c, d) iff there exists s ∈ S such that sad = sbc. We then define
a b at + bs
+ := ,
s t st
a b ab
· := .
s t st
We must check that these operations are well-defined on equivalence classes; this is
left as a (perhaps somewhat tedious, but not difficult) exercise for the reader.
Exercise 7.2. Indeed, check that S −1 R is a ring and that x 7→ x1 defines a
homomorphism of rings R → S −1 R. Thus S −1 R is an R-algebra, and in particular
an R-module.
Exercise 7.3. Let R be a domain and S = R• = R \ {0}. Show: S −1 R is
indeed the fraction field of R.
When f ∈ R, we denote the localization of R at the multiplicative subset generated
by f as Rf .
Example 7.1. Suppose f ∈ R is a nilpotent element: f n = 0 for some n ∈ Z+ .
n−1
Then 1 = ff n−1 whereas 0 = f0 . Since (f n−1 · f − f n−1 · 0) = 0, we have that
1 = 0, i.e., Rf is the zero ring. Conversely, if f is not nilpotent, then if it is a
unit, Rf = R. Otherwise, all the powers of f are distinct, and then f1 6= 11 , since
for any n ∈ N, f n (1 − f ) 6= 0, so Rf is not the zero ring. In general, S −1 R is the
zero ring iff S contains 0. This is to be regarded as a trivial case, and may safely
be tacitly excluded in the sequel.

Exercise 7.4. a) Show: the kernel of the natural map R → S −1 (R) is the set
of all r ∈ R such that for some s ∈ S, sr = 0.
b) The map R → S −1 (R) is injective iff S has no zerodivisors.
c) Show that the subset Q of all nonzerodivisors of a ring R is multiplicatively
2. PUSHING AND PULLING VIA A LOCALIZATION MAP 147

closed. The localization Q−1 R is called the total fraction ring of R. Show that
Q−1 (R) is a field iff R is a domain.
Exercise 7.5. Show: the homomorphism R → S −1 R is universal for homo-
morphisms R → T with f (S) ⊂ T × .
Exercise 7.6. A multiplicatively closed subset S of a ring R is saturated if
for all s ∈ S and all t ∈ R, if t | s then t ∈ S.
a) For a multiplicatively closed subset S ⊂ R, define its saturation S = {t ∈ R | t |
s for some s ∈ S}. Show that S contains S, is multiplicatively closed and saturated,
and is minimal with these properties: if T ⊃ S is saturated and multiplicatively
closed, then T ⊃ S.
b) Show: S−1 R and S −1 R are canonically isomorphic.
c) Let I be an ideal of R. Show that the following are equivalent:
(i) R \ I is multiplicatively closed.
(ii) R \ I is multiplicatively closed and saturated.
(iii) I is a prime ideal.
d) Let S ⊂ R be a saturatedT multiplicatively closed subset. Show: there is a subset
Y ⊂ Spec R such that S = p∈Y (R \ p).
(Hint: use Multiplicative Avoidance.)

2. Pushing and pulling via a localization map


Let R be a ring and S a multiplicatively closed subset. Let ι : R → S −1 R be the
natural map. As for any homomorphism of rings, ι induces maps between the sets
of ideals of R and the set of ideals of S −1 R, in both directions:
ι∗ : IR → IS −1 R , I 7→ IS −1 R,
ι∗ : IS −1 R → IR , J 7→ ι−1 (J).
Lemma 7.2. Let ι : R → S −1 R be as above. For an ideal I of R, we have
x
ι∗ (I) = { ∈ S −1 R | x ∈ I, s ∈ S}.
s
Proof. Let us temporarily write
x
I = { ∈ S −1 R | x ∈ I, s ∈ S}.
s
We want to show that I = ι∗ (I) = hι(I)iS −1 R . It is clear that ι(I) ⊂ I ⊂ ι∗ (I), so
it is enough to show that I is itself an ideal of S −1 R. No problem: if xs11 , xs22 ∈ I,
x1 x2 x1 s2 + x2 s1
+ = ∈ I,
s1 s2 s1 s2
y
and if s ∈ S −1 R, then
y x1 x1 y
= ∈ I. 
s s1 ss1
Like quotient maps, any localization map has the pull-push property.
Proposition 7.3. Let ι : R → S −1 R be as above. For an ideal J of S −1 R, we
have
J = ι∗ ι∗ J.
148 7. LOCALIZATION

Proof. We have seen before that for any homomorphism ι : R → R0 of rings


and any ideal J of R0 we have
J := ι∗ ι∗ J ⊂ J.
Thus it is enough to show the reverse containment. For this, consider an arbitrary
element xs ∈ J. Then x = s xs ∈ J hence also x ∈ ι∗ (J), so ι(x) ∈ J. But since J is
an ideal and s is a unit in S −1 R, we then also have 1s x = xs ∈ J. 

Lemma 7.4. Let ι : R → S −1 R be as above and I an ideal of R. The following


are equivalent:
(i) I ∩ S 6= ∅.
(ii) ι∗ (I) = S −1 R.
Proof. (i) =⇒ (ii): If s ∈ S ∩ I, then s ∈ IS −1 R, so 1 = ss ∈ ι∗ (I).
(ii) =⇒ (i): Suppose 1 ∈ ι∗ (I). By Lemma 7.2, 11 = xs for some x ∈ I and s ∈ S.
Clearing denominators, there is s0 ∈ S such that ss0 = s0 x and thus ss0 ∈ I ∩ S. 
Proposition 7.5. Let ι : R → S −1 R be a localization homomorphism.
a) For a prime ideal p of R, the following are equivalent:
(i) The pushforward ι∗ p is prime in S −1 R.
(ii) The pushforward ι∗ p is proper in S −1 R.
(iii) We have p ∩ S = ∅.
b) If p is prime and disjoint from S, then ι∗ (ι∗ p) = p.
Proof. a) (i) =⇒ (ii) since prime ideals are proper.
(ii) ⇐⇒ (iii) for all ideals of R by Lemma 7.4.
(iii) =⇒ (i): Suppose p is a prime ideal of R, and suppose we have as11 , as22 ∈ S −1 R
with as11 as22 = xs ∈ ι∗ (p). Clearing denominators, there is s0 ∈ S such that
ss0 a1 a2 = s0 s1 s2 x ∈ p.
Since S ∩ p = ∅, (ss0 ) ∈
/ p, and since p is prime, we conclude that a1 a2 ∈ p and
then that ai ∈ p for some i, hence asii ∈ ι∗ p for some i and ι∗ (p) is prime. This
completes the proof of part a).
b) Recall: for any homomorphism ι : R → R0 and any ideal I of R we have
ι∗ (ι∗ (I)) ⊃ I,
so taking I = p to be prime it suffices to show the inverse inclusion. Suppose
x ∈ ι∗ ι∗ p, i.e., there exist a ∈ p, s ∈ S such that ι(x) = x1 = as . By definition, this
means that there exists some s0 ∈ S such that s0 sx = s0 a ∈ p. Therefore either
s0 s ∈ p or x ∈ p, but since s0 s ∈ S and S is disjoint from p, we must have x ∈ p. 
Corollary 7.6. The maps ι∗ and ι∗ give mutually inverse bijections from the
set of prime ideals of S −1 R to the set of prime ideals of R that are disjoint from S.
Therefore we may – and shall – view Spec S −1 R as a subset of Spec R.
Exercise 7.7. a) Show: the results of Proposition 7.5 extend to all primary
ideals of R.1

1Recall p is primary if for a, b ∈ R such that ab ∈ p, either a ∈ p or bn ∈ p for some n ∈ Z+ .


We have not yet done much with this concept, and will not really address it squarely until the
section on primary decomposition.
4. COMMUTATIVITY OF LOCALIZATION AND PASSAGE TO A QUOTIENT 149

b) Let I be any ideal of R. Show that

ι∗ ι∗ I = {x ∈ R | ∃s ∈ S such that sx ∈ I}.

c) Exhibit a map ι : R → S −1 R and a (nonprimary) ideal I of R such that ι∗ ι∗ I ) I.

3. The fibers of a morphism


Let f : R → S be a homomorphism of rings, and let p ∈ Spec R. Consider the
“fiber of f ∗ : Spec S → Spec R over p”, i.e.,

fp = (f ∗ )−1 (p) = {P ∈ Spec S | f ∗ (P) = p}.

We claim that fp is canonically isomorphic to the spectrum of a certain ring.


Namely, let k(p) be the fraction field of the domain R/p. Then we wish to identify
fp with Spec(S ⊗R k(p)).

Let ι1 : S → S ⊗R k(p) and ι2 : k(p) → S ⊗R k(p) be the canonical maps. The ten-
sor product of R-algebras fits into a commutative square (INSERT) and is indeed
the categorical pushout: in other words, given any ring A and homomorphisms
ϕ1 : A → S ϕ2 : A → k(p) such that the composite homomorphisms ι1 ◦ϕ1 = ι2 ◦ϕ2
are equal, there exists a unique homomorphism Φ : A → R such that f ◦ Φ = ϕ1
and q ◦ Φ = ϕ2 , where q : R → R/p is the quotient map.

On the spectral side, all the arrows reverse, and the corresponding diagram is
(INSERT), which expresses Spec(S ⊗R k(p)) as the fiber product of Spec S and
Spec k(p) over Spec R.

Observe that the map ι1 : S → S ⊗R k(p) is the composite of the surjective map
q1 : S → S ⊗R R/p with the map `2 : S ⊗R R/p → (S ⊗R R/p) ⊗R/p k(p), the latter
map being localization with respect to the multiplicatively closed subset q1 (R \ p).
Both q1∗ and `∗2 are injections, and therefore ι1 ∗ = q1∗ ◦ `∗2 is injective. Similarly
Spec k(p) ,→ Spec R (this is just the special case of the above with R = S). It
follows that the above diagram identifies Spec S ⊗R k(p) with the prime ideals P
of Spec S such that f ∗ P = p.

4. Commutativity of localization and passage to a quotient


Lemma 7.7. Let R be a ring, S ⊂ R a multiplicatively closed subset, and I an
ideal of R. Write q : R → R/I for the quotient map and put S := q(S). Then there
is a canonical isomorphism
−1
S −1 R/IS −1 R ∼
= S (R/I).

Proof. Explicitly, we send as (mod I)S −1 R to as , where a = a + I, s = s + I.


It is straightforward to check that this an isomorphism. 

Matsumura makes the following nice comment: both sides satisfy the universal
property for homomorphisms f : R → R0 such that f (S) ⊂ (R0 )× and f (I) = 0.
Therefore they must be canonically isomorphic.
150 7. LOCALIZATION

5. Localization at a prime ideal


An extremely important example of a multiplicative subset of R is the complement
R \ p of a prime ideal p. As a matter of notation, we write Rp for (R \ p)−1 R.2
Proposition 7.8. If p is a prime ideal of R, then the localization Rp is a local
ring with unique maximal ideal pRp .
Proof. We know that the primes of the localized ring are precisely the push-
forwards of the prime ideals of R which are disjoint from the muliplicatively closed
set. Here S = R \ p, so being disjoint from S is equivalent to being contained in p.
Thus the unique maximal such element is indeed pRp . 
Remark: We will simply write p for the maximal ideal pRp of Rp .

Proposition 7.8, simple though it is, is of inestimable importance. It shows that


the effect of localization at a prime ideal on the lattice of ideals is dual to that of
passage to the quotient: if we mod out by a prime p, we get a ring R/p whose ideals
are precisely the ideals of R containing p. However, if we localize at R \ p, we get
a ring whose ideals are precisely the ideals of R contained in p. In particular, this
construction motivates us to develop an especially detailed theory of local rings, by
assuring us that such a theory could be put to good use in the general case.
Exercise 7.8. True or false: If (R, m) is a local ring and S ⊂ R• is a multi-
plicatively closed set, then S −1 R is a local ring (or the zero ring).

6. Localization of modules
If S is any multiplicative subset and M is any R-module, we can also construct
a localized R-module S −1 M . One the one hand, we can construct this exactly as
we did S −1 R, by considering the appropriate equivalence relation on pairs (m, s) ∈
M × S. On the other hand, we can just take the base extension S −1 R ⊗ M . We
are left with the task of showing that these two constructions are “the same”.
Exercise 7.9. Formulate a universal mapping property for the localization
morphism M → S −1 M . Check that both of the above constructions satisfy this
universal mapping property, and deduce that they are canonically isomorphic.
Exercise 7.10. a) Show: the kernel of M → S −1 M is the set of m ∈ M such
that ann(m) ∩ S 6= ∅.
b) Let R be a domain with fraction field K. Let M be an R-module. Show:
Ker(M → M ⊗ K) = M [tors].
c) Use part b) to give a new proof of Proposition 3.8b).
Exercise 7.11. Let N be any S −1 R-module. Show that there exists an R-
module M such that N ∼
= S −1 R ⊗R M .
Generally speaking, thinking of S −1 M as S −1 R⊗R M is more convenient for proving
results, because it allows us to employ the theory of tensor products of modules.
For example:
2This is inevitably a bit confusing at first, but our choice of notation for a loacalization is
designed to make this less confusing. The other common notation for the localization, RS , creates
a notational nightmare. As a mnemonic, remember that we gain nothing by localizing at a subset
S containing 0, since the corresponding localization is the trivial ring.
6. LOCALIZATION OF MODULES 151

Proposition 7.9. For any ring R and multiplicatively closed subset S of R,


S −1 R is a flat R-module. Equivalently, if
0 → M 0 → M → M 00 → 0
is a short exact sequence of R-modules, then
0 → S −1 M 0 → S −1 M → S −1 M 00 → 0
is a short exact sequence of R-modules (or equivalently, of S −1 R-modules).
Proof. Tensor products are always right exact, so we need only show S −1 M 0 ,→
0
S −1 M . Suppose not: then there exists m0 ∈ M 0 and s ∈ S such that ms = 0 ∈ M .
0
Thus there is g ∈ S such that gm0 = 0, but if so, then ms = 0 in M 0 .3 
Corollary 7.10. Let N and P be submodules of an R-module M . Then:
a) S −1 (N + P ) = S −1 N + S −1 P .
b) S −1 (N ∩ P ) = S −1 N ∩ S −1 P .
c) S −1 (M/N ) ∼=S −1 R S −1 M/S −1 N .
Exercise 7.12. Prove Corollary 7.10.
Proposition 7.11. Let M and N be R-modules and S a multiplicatively closed
subset of R. Then the mapping
m n m⊗n
⊗ 7→
s t st
−1
induces an isomorphism of S (R)-modules

S −1 M ⊗S −1 R S −1 N → S −1 (M ⊗R N ).
In particular, for any prime ideal p of R, we have

Mp ⊗Rp Np → (M ⊗R N )p .
Exercise 7.13. Prove Proposition 7.11.
Exercise 7.14. Let R be a ring, S ⊂ R multiplicative, and M an R-module.
a) If M is finitely generated, then S −1 M is a finitely generated S −1 R-module.
b) If M is finitely presented, then S −1 M is a finitely presented S −1 R-module.4
Recall that if M1 and M2 are R-submodules of an R-module M , then
(M1 : M2 ) = {x ∈ R | xM2 ⊂ M1 }
and thus
(M1 : M2 ) = ann((M1 + M2 )/M1 )).
Proposition 7.12. Let S ⊂ R be a multiplicatively closed subset.
a) Let M be a finitely generated R-module. Then
S −1 ann M = ann S −1 M.
b) If M1 , M2 are submodules of an R-module M and M2 is finitely generated, then
S −1 (M1 : M2 ) = (S −1 M1 : S −1 M2 ).
3Note also that the exactness of a sequence of R-modules does not depend on the R-module
structure but only on the underlying commutative group structure. Thus if we have a sequence
of commutative groups which can be viewed as a sequence of R-modules and also as a sequence
of R0 -modules, then exactness as R-modules is equivalent to exactness as R0 -modules.
4Actually both parts hold for any base change R → R0 ! We record it in this form since it
will be used later.
152 7. LOCALIZATION

Proof. a) We go by induction on the number n of generators of M .


Base Case: If n = 1 then M ∼
= R/I for some ideal I, so
S −1 ann M = S −1 I = ann S −1 R/S −1 I = ann S −1 M.
Induction Step: Suppose n ≥ 2 and that the result holds for all modules that can
be generated by n elements. Write M = M1 +M2 with M1 generated by n elements
and M2 generated by 1 element. Then
S −1 ann M = S −1 ann(M1 +M2 ) = S −1 (ann M1 ∩ann M2 ) = S −1 ann M1 ∩S −1 ann M2
= ann S −1 M1 ∩ ann S −1 M2 = ann S −1 M1 + S −1 M2 = ann S −1 M.
b) Put M := (M1 + M2 )/M1 . Then M is a quotient of M2 hence finitely generated,
so by part a) we have
S −1 (M1 : M2 ) = S −1 ann(M1 + M2 )/M1 = ann S −1 (M1 + M2 )/M1
= ann(S −1 (M1 + M2 )/S −1 M1 ) = (S −1 M1 : S −1 M2 ). 

7. Local properties
We say that a property P of a ring R is localizable if whenever R satisfies property
P , so does Rp for every prime ideal p of R. We say that a property P is local-
to-global if whenever Rp has property P for all prime ideals p of R, then R has
that property. Finally, we say a property is local if it is both localizable and
local-to-global. There are similar definitions for properties of R-modules.
Exercise 7.15.
a) Show the following properties of rings are localizable: being a field, having char-
acteristic 0, having prime characteristic p, being a domain, being reduced.
b) Show that the following properties of modules are localizable: freeness, projectiv-
ity, flatness, cyclicity, finite generation, finite presentation.
Exercise 7.16. Insert your own exercise here.
For a property P of rings, we say that a ring R is locally P if for all p ∈ Spec R,
Rp has the property P . Similarly, if Q is a property of modules, we say that an
R-module M is locally P if for all p ∈ Spec R, Mp has property Q.

Warning: It would also be reasonable to define “locally P” to mean that for


all p ∈ Spec R, there is f ∈ R \ p such that Rf (or Mf ) has property P . In the case
of rings, as we shall see later this latter definition means that Spec R has property
P locally with respect to the Zariski topology. We will consider this property as
well, but we call it Z-locally P instead.
Exercise 7.17. Let P be a property of rings (or modules). Show: “locally P ”
is a local property.
One of the most important themes in commutative algebra is the recognition of the
importance of local properties for rings and modules.

Remark: Very often it is true that if P is a local property, then R has prop-
erty P iff Rm has property P for all maximal ideals m of R. We will not introduce
terminology for this, but watch for it in the upcoming results.

First of all, for an R-module, being zero is a local property.


7. LOCAL PROPERTIES 153

Proposition 7.13. For an R-module M , the following are equivalent:


(i) M = 0.
(ii) Mp = 0 for all primes p of R.
(iii) Mm = 0 for all maximal ideals m of R.
Proof. Clearly (i) =⇒ (ii) =⇒ (iii), so asume that Mm = 0 for all maximal
ideals m of R. Suppose there exists 0 6= x ∈ M , and let I be the annihilator of x, so
that I is a proper ideal of R and thus contained in some maximal ideal m. Then x
is not killed by any element of the multiplicative subset R \ m and therefore maps
to a nonzero element of Mm : contradiction. 
Proposition 7.14. Let f : M → N be an R-module homomorphism.
a) The following are equivalent:
(i) f is injective.
(ii) For all prime ideals p of R, fp : Mp → Np is injective.
(iii) For all maximal ideals m of R, fm : Mm → Nm is injective.
b) Part a) holds with “injective” replaced everywhere by “surjective”, and thus also
if “injective” is replaced everywhere by “is an isomorphism.”
Proof. a) (i) =⇒ (ii) by the exactness of localization, and obviously (ii)
=⇒ (iii). Assume (iii), and let M 0 = Ker(f ). Then 0 → M 0 → M → N is exact,
0
hence for all m we have 0 → Mm → Mm → Nm is exact. So, by our assumption,
0
Mm = 0 for all maximal ideals m, and thus by Proposition 7.13 we have M 0 = 0.
The proof of part b) is virtually identical and left to the reader. 
Warning: Proposition 7.14 does not say: if M and N are R-modules such that
Mp ∼= Np as Rp modules for all p ∈ Spec R, then M ∼
= N . This is being asserted
only when there is a map f : M → N inducing all the isomorphisms between
localized modules.
Exercise 7.18. Exhibit finitely generated R-modules M and N which are “lo-
cally isomorphic” – i.e., Mp ∼
= Np for all p ∈ Spec R – but are not isomorphic.5
Corollary 7.15. Let R be a domain with fraction field K. Then we have
\
Rm = R.
m∈MaxSpec R
T
Proof. Certainly R ,→ m∈MaxSpec R Rm . Conversely, if x ∈ K \ R then
I := (R : Rx)
is a proper ideal, hence contained in some maximal ideal m. Because Rx is a finitely
generated R-module, by Proposition 7.12 we have
(Rm : Rm x) = Im ( Rm .
Thus x ∈
/ Rm . 
Let R be a domain with fraction field K, and let V be a finite-dimensional R-vector
space. An R-lattice in V is a finitely generated R-submodule Λ of V such that
hΛiK = V . Since K is a torsionfree R-module, every R-lattice in K is a finitely
generated, torsionfree R-module. Convesrsely, if Λ is a finitely generated torsionfree
R-module, let V := Λ ⊗R K. Then by Exercise 3.44 the natural R-module map
5In mantra form: “being isomorphic” is not a local property, but “being an isomorphism” is.
154 7. LOCALIZATION

Λ ,→ Λ ⊗R K is injective and makes Λ into an R-lattice in the finite-dimensional


K-vector space Λ ⊗R K. Such lattices are ubiquitous in algebra and number theory.
Corollary 7.15 extends to a local-global principle for lattices.
Theorem 7.16. Let R be a domain with fraction field K, let V be a finite-
dimensional K-vector space, and let Λ be a finitely generated R-submodule of K.
Then, inside V we have \
Λm = Λ.
m∈MaxSpec R
T
Proof. Put L := m∈MaxSpec R Λm . Clearly Λ ,→ L. Seeking a contradiction,
suppose there is x ∈ L \ Λ, and put Λ̃ = hΛ, xi. Then
I := (Λ : Λ̃)
is a proper ideal of R, hence contained in a maximal ideal m. Since Λ is finitely
generated, so is Λ̃, so by Proposition 7.12 we have
(Λm : Λ̃m ) = Im ( Rm .
It follows that Λ̃m = hΛ, xiRm ) Λm so x ∈
/ Λm , hence x ∈
/ L: contradiction. 
We give an application in the theory of stably free modules.
Proposition 7.17. A stably free module of rank one is free.
Proof. The natural proof uses exterior products of modules, which we have
unfortunately not defined in these notes. For the basics here see BOURBAKI or
[Ei, Appendix A2]. Especially, all the properties of exterior powers that we use
appear in [Ei, Prop. A2.2].
Now suppose P is such that P ⊕ Rn−1 ∼ = Rn . Taking top exterior powers we
get
n n i j
R∼ Rn ∼= (P ⊕ Rn−1 ) ∼
^ ^ M ^ ^
= = P⊗ Rn−1
i+j=n
2 n−2
!

^ ^
n−1
=M⊕ M⊗ R ⊕ ....
Vi Vi
For any p ∈ Spec R, Mp is free of rank one over Rp . Thus M p = ( M )p = 0
M = 0 for all i ≥ 2, so R ∼
Vi
for all i ≥ 2. By Proposition 7.13, = M. 
One of the most important local properties of an R-module M is the condition
of being locally free: for all p ∈ Spec M , Mp is free. We had better repeat the
previous warning in this special case: some people say that M is locally free iff for
all p ∈ Spec R there is f ∈ R \ p such that Mf is free. This is in general a stronger
property, which we call Z-locally free.

In light of the fact the being projective is a localizable property, Theorem 3.73
can be rephrased as follows.
Theorem 7.18. (Kaplansky) A projective module is locally free.
From our study of vector bundles it is natural to wonder about the converse: must
a finitely generated locally free module be projective? In complete generality the
answer is negative (we will meet counterexamples later on), but morally it is right:
cf. Theorem 7.29.
7. LOCAL PROPERTIES 155

7.1. Rank Functions.

Let M be a finitely generated module. There is an associated rank function


rankM : Spec R → N, given by
rankM (p) = dimk(p) M ⊗ k(p).
Here k(p) is the fraction field of the domain R/p, and since M is a finitely generated
R-module, M ⊗ k(p) is a finite-dimensional k(p)-vector space.
Exercise 7.19. a) Show: for finitely generated R-modules M and N , we have
rankM ⊕N = rankM + rankN
and
rankM ⊗R N = rankM · rankN .
b) Compute the rank function on the Z-module Z/nZ.
c) Compute the rank function on any finitely generated module over a PID.
Exercise 7.20. Let M be a finitely generated R-module.
a) Show: if M is locally free, then for all p ∈ Spec R, Mp ∼
rank (p)
= Rp M .
b) Suppose M is stably free of rank n: i.e., there are m, n ∈ N such that M ⊕ Rm ∼
=
Rm+n . Show: for all p ∈ Spec R, we have rankM (p) = n.
We will mostly be interested in the rank function on a finitely generated projective
module. As in §6, we view a finitely generated projective module as being anal-
ogous to a vector bundle, and then the rank at p plays the role of the dimension
of the fiber at p. In the case of a vector bundle on a topological space X, the
rank function is locally constant, hence constant if X is connected. Once we study
the Zariski topology on Spec R in earnest, we can and will prove the analogous
statement: the rank function on a finitely generated projective module M is locally
constant. Its failure to be constant is somehow the coarsest possible obstruction to
M being free: if the rank function is not constant, then M cannot even be stably
free: equivalently, its class in the reduced K-group K^0 (R) is nontrivial.

Here is a (not so deep) criterion for a projective module to be free.


Proposition 7.19. Let M be a projective R-module of constant rank n. Then
M is free iff M can be generated by n elements.
Proof. If M is free then M ∼ = Rn . Conversely, that M can be generated by n
elements means there is a surjective R-module map ϕ : Rn → M . Let K = ker ϕ;
since M is projective, we have
Rn ∼= K ⊕ M.
Thus K is also finitely generated projective, so for all p ∈ Spec R we have
Rn ∼
p= K p ⊕ Mp
with Kp ∼= Rp and Mp ∼
r(p)
= Rpn . It follows (by tensoring to k(p)) that Kp = 0.
Being zero is a local property, so K = 0 and ϕ : Rn → M is an isomorphism. 
We can now give a situation in which our coarsest possible obstruction to freeness
is the only one.
156 7. LOCALIZATION

Lemma 7.20. Let M be a finitely generated projective R-module, and let I be


an ideal contained in the Jacobson radical. If M/IM is free, then so is M .
Proof. If M/IM ∼ = (R/I)n , then by Nakayama’s Lemma n is the least number
of generators of M . Moreover, for all m ∈ MaxSpec R, since the map M → M/mM
factors through M/IM we have rankM (m) = n. Now observe that if p1 ⊂ p2 ∈
Spec R, then
rk
Rp1M
(p1 ) ∼
= Mp1 ∼
= Mp2 ⊗Rp2 Rp1 ∼
= Rp2M 2 ⊗Rp2 Rp1 ∼
rk (p ) rk (p )
= Rp1M 2 ,
so rkM (p1 ) = rkM (p2 ). It follows that M has constant rank n and can be generated
by n elements, so by Proposition 7.19 we have that M is free. 
Exercise 7.21. Let I ⊂ J(R) be an ideal of R. Let P1 , P2 be two finitely
generated projective R-modules. Show: if P1 /IP1 ∼ =R/I P2 /IP2 then P1 ∼
= P2 .

(Hint: use the projectivity of P1 to factor the map P1 → P1 /IP1 → P2 /IP2 through
P2 .)
Corollary 7.21. Let R be a semilocal ring (MaxSpec R is finite). Let M be
a finitely generated projective R-module. Then M is free iff it has constant rank.
Proof. We know that free modules have constant rank. Conversely, suppose
M is projective of constant rank n, and let m1 , . . . , mN be the distinct maximal
TN QN
ideals of R, so J(R) = i=1 mi = i=1 mi . By the Chinese Remainder Theorem
for modules we have
N
M/J(R) ∼
Y
= M/mi M.
i=1
Since M has constant rank n, we have dimR/mi M mi M = n for all i, so M/J(R)M
is a free R/J(R)-module of rank n. Apply Lemma 7.20. 
The topological analogue of Corollary 7.21 is, roughly, that a vector bundle on a
discrete space is trivial iff it has constant rank. (However that is a triviality whereas
Corollary 7.21 is actually rather useful.)

7.2. Local nature of flatness.


Proposition 7.22. For an R-module M , the following are equivalent:
(i) M is flat.
(ii) For all prime ideals p of R, Mp is flat.
(iii) For all maximal ideals m of R, Mm is flat.
Proof. (i) =⇒ (ii) is a special case of Proposition 7.9; (ii) =⇒ (iii) is imme-
diate. So assume (iii), and let N ,→ P be any injective R-module homomorphism.
Then, by exactness of localization, for all maximal ideals m we have Nm ,→ Pm .
Since Mm is assumed to be flat, we have (N ⊗R M )m = Nm ⊗Rm Mm ,→ Pm ⊗Rm
Mm = (P ⊗R M )m . Applying Proposition 7.14 we conclude that N ⊗R M → P ⊗R P
is injective, and therefore M is flat over R. 
Corollary 7.23. Let R be a ring, S ⊂ R a multiplicative subset. If M is a
flat R-module, then S −1 M is a flat S −1 R-module.
Proof. If M is flat, so is Mp for each prime ideal p of M , but since the primes
of S −1 R are a subset of the primes of R, this implies that S −1 M is flat. 
8. LOCAL CHARACTERIZATION OF FINITELY GENERATED PROJECTIVE MODULES157

When a property P of rings or modules is not local, it is often of interest to study


also its “localized version”: we say that an R-module M is locally P if for all
prime ideals p of R, Mp has property p (and similarly for rings).
7.3. Absolute flatness revisited.
Lemma 7.24. Suppose an absolutely flat ring R is either local or a domain.
Then R is a field.
Proof. Suppose R is not a field, and let x ∈ R be a nonzero, nonunit. Then
(0) ( (x) ( R. Proposition 3.100 gives R ∼
= (x) ⊕ J: contradiction. 
Lemma 7.25. Let R be a ring.
a) If R is absolutely flat and S ⊂ R is any multiplicative subset, then S −1 R is
absolutely flat.
b) R is absolutely flat iff for every maximal ideal m of R, Rm is a field.
Proof. a) By Exercise 7.11 every S −1 R-module is of the form S −1 R ⊗R M
for some R-module M . By hypothesis M is flat, so by Corollary 7.23, so is S −1 M .
b) If R is absolutely flat, and m is a maximal ideal of R, then by part a) Rm is
absolutely flat. On the other hand it is a local ring, so by Lemma 7.24, Rm is a
field. Conversely, assume that each Rm is a field, and let M be an R-module. Then
for all m ∈ MaxSpec R, Mm is a flat Rm -module, so M is a flat R-module. 
Theorem 7.26. For a ring R, the following are equivalent:
(i) R/ nil R is absolutely flat, i.e., every R/ nil R-module is flat.
(ii) Every prime ideal of R is maximal.
Proof. Since the prime ideals of R are the same as those of R/ nil R, it is
equivalent to prove the following simpler assertion: if R is reduced, it is absolutely
flat if and only if every prime ideal of R is maximal. Suppose R is absolutely flat
and p ∈ Spec R. Then R/p is an absolutely flat domain, hence a field by Lemma
7.24, hence p is maximal. Let m be a maximal ideal of R. Then Rm is a reduced
local ring, hence a field. By Lemma 7.25, R is absolutely flat. 

8. Local characterization of finitely generated projective modules


Let us call a family of {fi }i∈I of elements of R a Z-family if hfi i = 1. Clearly for
every Z-family there is a finite subset J ⊂ I such that {fi }i∈J is also a Z-family.
(Later on, this trivial observation will be dressed up in rather fancy attire: this
gives the quasi-compactness of the Zariski topology on Spec R.)

A property P of rings or modules will be said to be Z-local if it holds over R


iff it holds over all Rfi for some Z-family {fi } of R.
Proposition 7.27.
Let u : M → N be a homomorphism of R-modules, and let p ∈ Spec R.
a) If N is finitely generated and up is surjective, there exists f ∈ R \ p such that
uf : Mf → Nf is surjective.
b) The surjectivity of u is a Z-local property.
c) If M is finitely generated, N is finitely presented and up is an isomorphism, then
there exists f ∈ R \ p such that uf : Mf → Nf is an isomorphism.
d) If M is finitely generated and N is finitely presented, then the bijectivity of u is
a Z-local property.
158 7. LOCALIZATION

Proof. Write out the exact sequence


u
0 → ker u → M → N → coker u → 0.
By the flatness of localization, this sequence remains exact upon being tensored
with Rf for any f ∈ R or with Rp for any p ∈ R. It follows that passage to the
kernel and cokernel commutes with localization.
a) We’re assuming 0 = coker(up ) = (coker u)p , i.e., for each x ∈ coker u there exists
fx ∈ R \ p such that fx x = 0. Since coker u is a quotient of the finitely generated
module N , it is finitely generated, say by x1 , . . . , xn . Then f = fx1 · · · fxn ∈ R \ p
is such that f coker u = 0, so 0 = (coker u)f = coker(uf ) and uf is surjective.
b) It is clear that if u is surjective, then for any f ∈ R, uf is surjective. Conversely,
let {fi }i∈I be a Z-family such that ufi is surjective for all i. Then for any p ∈ Spec R
there exists i ∈ I such that fi ∈ R \ p, so that up is a further localization of ufi and
thus the surjectivity of ufi implies that of up . By Proposition 7.14, u is surjective.
c) By part a), there exists f1 ∈ R \ p such that coker uf1 = 0, and thus we have an
exact sequence
0 → (ker u)f1 → Mf1 → Nf1 → 0.
Since N is finitely presented over R, Nf1 is finitely presented over Rf1 and thus
(ker u)f1 is finitely generated. Arguing as in part b), we get f2 ∈ R \ p such that

f1 f2 ker u = 0. Taking f = f1 f2 we get uf : Mf → Nf .
d) This is proved analogously to part b) and is left to the reader. 
Corollary 7.28. For a finitely presented R-module M , the following are equiv-
alent:
(i) There is a Z-family {fi }i∈I of R such that for all i ∈ I, Mfi is a free Rfi -module.
(ii) M is locally free: for all p ∈ Spec R, Mp is a free Rp -module.
(ii0 ) For every m ∈ MaxSpec R, Mm is a free Rm -module.
Proof. (i) =⇒ (ii): For each prime ideal p there exists at least one i such
that fi ∈
/ p; equivalently, the multiplicative subset generated by fi is contained in
R \ p. Thus Mp = Mfi ⊗Rfi Rp and since Mfi is free, so is Mp .
(ii) =⇒ (i): It is enough to find for each prime ideal p an element fp ∈ R \ p such
that Mfp is free: for if so, then {fp }p∈Spec R is a Z-family. Choose x1 , . . . , xn ∈ M
whose images in Mp give an Rp -basis, and define u : Rn → M via ei 7→ xi . Then
up is an isomorphism, so by Proposition 7.27c) we may choose fp ∈ R \ p such that
ufp is an isomorphism and thus Mfp is free.
(ii) ⇐⇒ (ii0 ): Since freeness is localizable, this follows from Exercise 7.16. 
Remark 7. (M. Brandenburg) The proof of (ii) =⇒ (i) shows something a bit
stronger: if M is a finitely presented module over a ring R and for some p ∈ Spec R
we have that Mp is free, then there is fp ∈ R such that Mfp is free.

Qn For 1 ≤ i ≤ n, let Mi be a finitely Q


Exercise 7.22. generated proejctive Ri -
n
module. Show: i=1 Mi is a finitely generated projective i=1 Ri -module.
We can now prove one of the major results of this text.
Theorem 7.29. For an R-module M , the following are equivalent:
(i) M is finitely generated and projective.
(ii) M is finitely presented and locally free.
(iii) For every maximal ideal m of R, there exists f ∈ R \ m such that Mf is a
8. LOCAL CHARACTERIZATION OF FINITELY GENERATED PROJECTIVE MODULES159

locally free Rf -module of finite rank.


(iv) There exists a finite Z-family {f1 , . . . , fn } of R such that hf1 , . . . , fn i = R and
for all i, Mfi is a finitely generated free Rfi -module.
Proof. (i) =⇒ (ii): Let M be finitely generated and projective. There
exists a finitely generated free module F and a surjection q : F → M . Since M is
projective, q splits and Ker(q) is not just a submodule of F but also a quotient and
thus finitely generated. So M is finitely presented. Since projectivity is preserved
by base change and any finitely generated projective module over a local ring is
free (Theorem 3.16), for all maximal ideals m of R, Mm is free.
(ii) =⇒ (iii): this follows immediately from Corollary 7.28.
(iii) =⇒ (iv): For each m ∈ MaxSpec R, choose fm ∈ R \ m such that Mfm is a
finitely generated free Rfm -module. Then {fm }m∈MaxSpec R is a Z-family of R, and
as remarked above, every QnZ-family contains a finite subfamily.
(iv) =⇒ (i): Put S = i=1 Rfi and let f : R → S be the natural map.
Step 1: First note that
n
\ n
\
ker f = ker(R → Rfi ) = ann(fi ) = f annhf1 , . . . , fn i = ann R = 0,
i=1 i=1

so f is injective, and thus S is an extension ring of R.


Step 2: We claim f : R ,→ S is a faithfully flat extension. Since localizations
are flat and direct sums of flat algebras are flat, S/R is a flat extension. So by

`n it is enough to ∗show that f : Spec S → Spec R is surjective. But
Theorem 3.108,
Spec S = i=1 Spec Rfi and f (Spec Rfi ) is the subset of p ∈ Spec R such that
fi ∈
/ p. Since {f1 , . . . , fn } forms a Z-family, no proper ideal can contain all the fi ’s,
and therefore p lies in at least one f ∗ (Spec Rfi ).
Step 3: We have aQfaithfully flat ring extension f : R ,→ S and an R-module M such
n Qn
that M ⊗R S = i=1 Mfi is a finitely generated projective S = i=1 Rfi -module
(Exercise 7.22). By Theorem 3.111, M is finitely generated and projective! 

Corollary 7.30. Every finitely presented flat R-module is projective.


Proof. Let M be a finitely presented, flat R-module. For each maximal ideal
m of R, Mm is a finitely presented flat module over the local ring Rm , hence is free
by Theorem 3.51. Therefore by criterion (iii) of Theorem 7.29, M is projective. 

Corollary 7.31. Let R be a Noetherian ring, and let M be a finitely generated


R-module. The following are equivalent:
(i) M is projective.
(ii) M is locally free.
(iii) M is flat.
Exercise 7.23. Prove Corollary 7.31.
Theorem 7.32. For an R-module A, the following are equivalent:
(i) A is finitely generated projective.
(ii) For all R-modules B, the natural map
Φ : A∨ ⊗R B → HomR (A, B)
induced by (f, b) 7→ (a 7→ f (a)b) is an isomorphism.
(iii) The map Φ : A∨ ⊗R A → HomR (A, A) is an isomorphism.
160 7. LOCALIZATION

Proof. (i) =⇒ (ii): It is enough to show that for all p ∈ Spec R, Φp is an


isomorphism. Since A is finitely generated projective, it is finitely presented; more-
over Rp is a flat R-module, so by Theorem 3.103 we have a canonical isomorphism
HomR (A, N ) ⊗R Rp = HomRp (Ap , Np ). Also tensor products commute with base
change, so it is enough to show
Φp : A ∨
p ⊗Rp Bp → HomRp (Ap , Bp )
is an isomorphism. Since A is finitely generated projective, Ap is finitely generated
and free. We are thus essentially reduced to a familiar fact from linear algebra,

namely the canonical isomorphism V ∨ ⊗ W → Hom(V, W ) for vector spaces over a
field, with V finite-dimensional. We leave the details to the reader as an exercise.
(ii) =⇒ (iii): This is immediate.P
m
(iii) P=⇒ (i): Let Φ−1 (1A ) = i=1 fi ⊗ ai . Then we have that for all a ∈ A,
m
a = i=1 fi (a)ai . By the Dual Basis Lemma, A is finitely generated projective. 
We end by showing that without the Noetherian hypothesis, a finitely generated
locally free module need not be projective.
Proposition 7.33. For a ring R, the following are equivalent:
(i) R is absolutely flat.
(ii) Every R-module is locally free.
Proof. (i) =⇒ (ii): By Lemma 7.25, for m ∈ MaxSpec R, Rm is a field, so
every Rm -module is free. By Theorem 7.26 every prime ideal of R is maximal, so
every R-module is locally free.
(ii) =⇒ (i): Applying Lemma 7.25 again, if R is not absolutely flat, there is
m ∈ MaxSpec R such that Rm is not a field, and thus there exists a nonfree Rm -
module Mm . By Exercise X.X, there is an R-module M such that M ⊗R Rm ∼ = Mm
and thus M is not locally free. 
Exercise 7.24. (G. Elencwajg)
a) Let I be an ideal of the ring R. Show: if R/I is projective, then I is principal.
b) Let R be a ring which is absolutely flat and not Noetherian (e.g. an infinite
product of fields). Let I be an ideal which is not finitely generated. Show: R/I is
finitely generated, flat, not projective, locally free and not Z-locally free.
CHAPTER 8

Noetherian rings

We have already encountered the notion of a Noetherian ring, i.e., a ring in which
each ideal is finitely generated; or equivalently, a ring which satisfies the ascending
chain condition (ACC) on ideals. Our results so far have given little clue as to the
importance of this notion. But in fact, as Emmy Noether showed, consideration of
rings satisfying (ACC) is a major unifying force in commutative algebra.

In this section we begin to see why this is the case. After giving an introduc-
tory examination of chain conditions on rings and modules, we are able to make
the key definitions of height of a prime ideal and dimension of a ring, which we
will slowly but surely work towards understanding throughout the rest of these
notes. Indeed we begin by giving a reasonably complete analysis of the structure
theory of Artinian rings, which, as we will show, really is our first order of busi-
ness in attempting the systematic study of Noetherian rings, since according to the
Akizuki-Hopkins theorem the Artinian rings are precisely the Noetherian rings of
dimension zero. We are then able to state and prove three of the most important
and useful theorems in the entire subject. Whereas the first theorem, the Hilbert
basis theorem, gives us a large supply of Noetherian rings, the latter two theorems,
Krull’s intersection theorem and Krull’s principal ideal theorem, are basic results
about the structure theory of Noetherian rings.

1. Chain conditions on partially ordered sets


Proposition 8.1.
For a partially ordered set (S, ≤), the following are equivalent:
(i) S satisfies the Ascending Chain Condition (ACC): there is no infinite se-
quence {xi }∞n=1 of elements of S with xn < xn+1 for all n ∈ Z .
+

(ii) Every nonempty subset T ⊂ S has a maximal element.


A partially ordered set satisfying these equivalent conditions is called Noetherian.
Proof. (i) =⇒ (ii): Let T be a nonempty subset of S without a maximal
element. Since T is nonempty, choose x1 ∈ T . Since T has no maximal elements,
choose x2 > x1 . Since T has no maximal elements, choose x3 > x2 . And so on: we
get an infinite strictly ascending chain in S.
(ii) =⇒ (i): Indeed, an infinite strictly ascending chain is a nonempty subset
without a maximal element. 
Similarly, we say that a partially ordered set satisfies the Descending Chain
Condition (DCC) –if there is no infinite sequence {yj }∞ j=1 of elements of S such
that yj > yj+1 for all j ∈ Z+ . As above, this holds iff every nonempty subset of
S has a minimal element, and a partially ordered set satisfying these equivalent
conditions is called Artinian.
161
162 8. NOETHERIAN RINGS

Every partially ordered set (S, ≤) has an order dual S ∨ : the underlying set is
S, and we put x ≤∨ y ⇐⇒ y ≤ x. Clearly S is Noetherian (resp. Artinian) iff
S ∨ is Artinian (resp. Noetherian). Thus at this level of abstraction we really have
one notion here, not two. Nevertheless in our applications to rings and modules
the two conditions remain quite distinct.

Examples: If S is finite it satisfies both ACC and DCC. With the usual order-
ings, the positive integers Z+ satisfy DCC but not ACC, the negative integers Z−
(or equivalently, Z+ with the opposite ordering) satisfy ACC but not DCC, and
the integers Z satisfy neither.
Exercise 8.1. Let S be a partially ordered set.
a) Show that S satisfies (ACC) (resp. (DCC)) iff there is no order embedding
Z+ ,→ S (resp. Z− ,→ S).
b) Suppose S is totally ordered. Show that S satisfies (DCC) iff it is well-ordered:
i.e., every nonempty subset has a minimal element.

2. Chain conditions on modules


Let R be a ring, and M a (left) R-module. It makes sense to speak of the (ACC)
and (DCC) for R-submodules of M . Indeed, we will call M a Noetherian module
if it satisfies (ACC) and an Artinian module if it satisfies (DCC).
Exercise 8.2. Show: an R-module M is Noetherian iff every R-submodule M 0
of M is finitely generated.
Example 8.2. As a Z-module, the integers Z are Noetherian but not Artinian.
Example 8.3. As a Z-module, the group of all p-power roots of unity in the
complex numbers – in other words, limn→∞ µpn – is Artinian but not Noetherian.
Every ring R is naturally an R-module, and the R-submodules of R are precisely the
ideals. Thus it makes sense to say whether R is a Noetherian or Artinian R-module,
and – thank goodness – this is visibly consistent with the previous terminology.
Exercise 8.3. Let M 0 ⊂ M be R-modules, and ϕ : M → M/M 0 be the quotient
map. If N1 and N2 are submodules of M such that N1 ⊂ N2 , N1 ∩ M 0 = N2 ∩ M 0
and ϕ(N1 ) = ϕ(N2 ), show that N1 = N2 .
Theorem 8.4. Let 0 → M 0 → M → M 00 → 0 be a short exact sequence
of R-modules. Then M is Noetherian (resp. Artinian) iff both M 0 and M 00 are
Noetherian (resp. Artinian).
Proof. We do the Noetherian case, leaving the similar Artinian case as an
exercise for the reader. First, since an infinite ascending chain in a submodule
or quotient module of M gives rise to an infinite ascending chain in M , if M is
Noetherian, both M 0 and M 00 are. Conversely, suppose N1 ( N2 ( . . . is an
infinite ascending chain of submodules of M . Consider the chain (Ni + M 0 )/M 0 in
M 00 = M/M 0 . By hypothesis, this chain eventually stabilizes, i.e., for sufficiently
large i and j, Ni + M 0 = Nj + M 0 . Similarly, by intersecting with M 0 we get that
for sufficiently large i and j Ni ∩ M 0 = Nj ∩ M 0 . Applying Exercise 8.3 we conclude
Ni = Nj for all sufficiently large i, j. 
3. SEMISIMPLE MODULES AND RINGS 163

A ring R is Noetherian if R is a Noetherian R-module. A ring R is Artinian if


R is an Artinian R-module.
Exercise 8.4. Let R be a ring.
a) Show that R is Noetherian iff every finitely generated R-module is Noetherian.
b) Show that R is Artinian iff every finitely generated R-module is Artinian.
c) Exhibit a ring R which is Noetherian but not Artinian.
d) Can you find a ring R which is Artinian but not Noetherian?1

3. Semisimple modules and rings


In this section we allow not necessarily commutative rings R. By a “module over
R” we mean a left R-module unless otherwise indicated.

A module M is simple if it is nonzero and has no proper, nonzero submodules.

This definition is of course made in analogy to that of a simple group, namely


a nontrivial group possessing no nontrivial proper normal subgroups. And indeed
many of the results in this and subsequent sections were first proved in the context
of groups. It is even possible to work in a single context that simultaneously gen-
eralizes the case of groups and modules (over a not necessarily commutative ring),
the key concept being that of groups with operators. For more on this perspec-
tive we invite the reader to consult any sufficiently thick all-purpose graduate level
algebra text, the gold standard here being [J1], [J2].
Exercise 8.5. (Schur’s Lemma): Let M be a simple R-module. Show:
EndR (M ) is a division ring.
Theorem 8.5. For an R-module M , the following are equivalent:
(i) M is a direct sum of simple submodules.
(ii) Every submodule of M is a direct summand.
(iii) M is a sum of simple submodules.
A modules satisfying these equivalent conditions is called semisimple.
L
Proof. (i) =⇒ (ii): Suppose
L M = i∈I Si , with each Si a simple submodule.
For each J ⊂ I, put MJ = i∈J Si . Now let N be an R-submodule of M . An easy
Zorn’s Lemma argument gives us a maximal subset J ⊂ I such that N ∩ MJ = 0.
For i ∈/ J we have (MJ ⊕ Si ) ∩ N 6= 0, so choose 0 6= x = y + z, x ∈ N , y ∈ MJ ,
z ∈ Si . Then z = x − y ∈ (Mj + N ) ∩ Si , and if z = 0, then x = y ∈ N ∩ Mj = 0,
contradiction. So (MJ ⊕ N ) ∩ Si 6= 0. Since Si is simple, this forces Si ⊂ MJ ⊕ N .
It follows that M = MJ ⊕ N .
(ii) =⇒ (i): First observe that the hypothesis on M necessarily passes to all
submodules of M . Next we claim that every nonzero submodule C ⊂ M contains
a simple module.
proof of claim: Choose 0 6= c ∈ C, and let D be a submodule of C which
is maximal with respect to not containing c. By the observation of the previous
paragraph, we may write C = D ⊕ E. Then E is simple. Indeed, suppose not and
let 0 ( F ( E. Then E = F ⊕ G so C = D ⊕ F ⊕ G. If both D ⊕ F and D ⊕ G
contained c, then c ∈ (D ⊕ F ) ∩ (D ⊕ G) = D, contradiction. So either D ⊕ F
or D ⊕ G is a strictly larger submodule of C than D which does not contain c,
1More on this later!
164 8. NOETHERIAN RINGS

contradiction. So E is simple, establishing our claim.


Now let N ⊂ M be maximal with respect to being a direct sum of simple
submodules, and write M = N ⊕ C. If C 6= 0, then by the claim C contains a
nonzero simple submodule, contradicting the maximality of N . Thus C = 0 and
M is a direct sum of simple submodules.
(i) =⇒ (iii) is immediate.
(iii) =⇒ (i): as above, by Zorn’s Lemma there exists a submodule N of M which is
maximal with respect to being a direct sum of simple submodules. We must show
N = M . If not, since M is assumed to be generated by its simple submodules,
there exists a simple submodule S ⊂ M which is not contained in N . But since S
is simple, it follows that S ∩ N = 0 and thus N ⊕ S is a strictly larger direct sum
of simple submodules: contradiction. 

Corollary 8.6. An R-module M has a unique maximal semisimple sub-


module, called the socle of M and written Soc M . Thus M is semisimple iff
M = Soc M .
Exercise 8.6. Prove Corollary 8.6.
Exercise 8.7. Let N ∈ Z+ . Compute the socle of the Z-module Z/N Z. Show
in particular that Z/N Z is semisimple iff N is squarefree.
A not-necessarily-commutative ring R is left semisimple if R is semisimple as a
left R-module.
Theorem 8.7. For a nonzero not necessarily commutative ring R, the following
are equivalent:
(i) R is left semisimple.
(ii) Every left ideal of R is a direct summand.
(iii) Every left ideal of R is an injective module.
(iv) All left R-modules are semisimple.
(v) All short exact sequences of left R-modules split.
(vi) All left R-modules are projective.
(vii) All left R-modules are injective.
Proof. We will show (i) ⇐⇒ (ii), (iv) ⇐⇒ (v) ⇐⇒ (vi) ⇐⇒ (vii) and
(ii) =⇒ (vii) =⇒ (iii) =⇒ (ii), which suffices.
(i) =⇒ (ii) follows immediately from Theorem 8.5.
(iv) ⇐⇒ (v) follows immediately from Theorem 8.5.
(v) ⇐⇒ (vi) and (v) ⇐⇒ (vii) are immediate from the definitions of projective
and injective modules.
(ii) =⇒ (vii): Let I be a left ideal of R and f : I → M an R-module map. By
π
hypothesis, there exists J such that I ⊕ J = R, so f extends to F : R = I ⊕ J →1
I → M . By Baer’s Criterion, M is injective.
(vii) =⇒ (iii) is immediate.
(iii) =⇒ (ii) is immediate from the definition of injective modules. 

Lemma 8.8. Let R be a ring and {Mj }j∈J be an indexed family of nonzero
R-modules. The following are equivalent:
(i) I is finite
L and each Mj is finitely generated.
(ii) M = j∈J Mj is finitely generated.
3. SEMISIMPLE MODULES AND RINGS 165

Proof. (i) =⇒ (ii) is left to the reader as an easy exercise.


(ii) =⇒ (i): Each Mj is isomorphic to a quotient of M , so if M is finitely generated,
so is Mj . Now let X = {x1 , . . . , xn } be a finite generating
P set for M , and for each
1 ≤ 1 ≤ n, let xij be the j-component of xi , so xi = j∈J xij . This sum is of
course finite, and therefore the set J 0 ⊂ JLof indices j such that xij 6= 0 for some
1 ≤ i ≤ n is finite. It follows that hXi ⊂ j∈J 0 Mj ( M , contradiction. 
Lemma 8.9.
Qn Let R1 , . . . , Rn be finitely many not necessarily commutative rings,
and put R = i=1 Ri . Then R is semisimple iff Ri is semisimple for all 1 ≤ i ≤ n.
Exercise 8.8. Prove Lemma 8.9.
We now quote the following basic result from noncommutative algebra.
Theorem 8.10. (Wedderburn-Artin) For a ring R, the following are equiva-
lent:
(i) R is semisimple as a left R-module (left semisimple).
(ii) R is semisimple as a right R-module (right semisimple).
(iii) There are N, n1 , . . . , nN ∈ Z+ and division rings D1 , . . . , DN such that
N
R∼
Y
= Mni (Di ).
i=1
Combining Theorems 8.7 and 8.10 gives us a tremendous amount of information.
First of all, a ring is left semisimple iff it is right semisimple, so we may as well
speak of semisimple rings. A ring is semisimple iff it is absolutely projective
iff it is absolutely injective.

Coming back to the commutative case, the Wedderburn-Artin theorem tells us


that the class of semisimple / absolutely projective / absolutely injective rings is
extremely restricted.
Corollary 8.11. A commutative ring is semisimple iff it is a finite product
of fields.
However it is significantly easier to give a proof of Wedderburn-Artin in the com-
mutative case, so we will give a direct proof of Corollary 8.11
Proof. Step -1: Officially speaking the theorem holds for the zero ring because
it is an empty product of fields. In any event, we may and shall assume henceforth
that our semisimple ring is nonzero.
Step 0: A field is a semisimple ring: e.g. every module over a field is free, hence
projective. By Lemma 8.9, a finite direct product L of fields is therefore semisimple.
Step 1: Let R be a semisimple ring, and let R = i∈I Mi be a direct sum decompo-
sition into simple R-modules. R is a finitely generated R-module, by Lemma 8.8 I is
finite, and we may identify it with {1, . . . , n} for some n ∈ Z+ : R = M1 ⊕ . . . ⊕ Mn .
Step 2: We may uniquely write 1 = e1 + . . . + en with ei ∈ Mi . Then for all i 6= j,
ei ej = 0, and this together with the identity 1 · 1 = 1 implies that e2i = ei for all
i. As usual for idempotent decompositions, this expresses R as a direct product of
the subrings Ri = Mi = ei R. Moreover, since Mi is a simple R-module, Ri has no
proper nonzero ideals, and thus it is a field, say ki . 
Exercise 8.9. Exhibit an absolutely flat commutative ring that is not semisim-
ple.
166 8. NOETHERIAN RINGS

4. Normal Series
If M is an R-module a normal series is a finite ascending chain of R-submodules
0 = M0 ( M1 ( . . . ( Mn = M . We say that n is the length of the series. (The
terminology is borrowed from group theory, in which one wants a finite ascending
chain of subgroups with each normal in the next. Of course there is no notion of
“normal submodule”, but we keep the group-theoretic terminology.)

There is an evident partial ordering on the set of normal series of a fixed R-module
0
M : one normal series {Mi }ni=0 is less than another normal series {Mj0 }nj=0 if for all
1 ≤ i ≤ n, Mi is equal to Mj0 for some (necessarily unique) j. Rather than saying
that {Mi } ≤ {Mj0 }, it is traditional to say that the larger series {Mj0 } refines the
smaller series {Mi }.

Given any normal series {Mi } we may form the associated factor sequence
0
M1 /M0 = M1 , M2 /M1 , . . . , Mn /Mn−1 = M/Mn−1 . Two normal series {Mi }ni=0 , {Mj0 }nj=0
are equivalent if n = n0 and there is a permutation σ of {1, . . . , n} such that for all
0 0
1 ≤ i ≤ n, the factors Mi /Mi−1 and Mσ(i) /Mσ(i)−1 are isomorphic. In other words,
if we think of the factor sequence of a normal series as a multiset of isomorphism
classes of modules, then two normal series are equivalent if the associated multisets
of factors are equal.
Exercise 8.10. Show: refinement descends to a partial ordering on equivalence
classes of normal series of a fixed R-module M .
The following theorem is the basic result in this area.
Theorem 8.12. (Schreier Refinement) For any R-module M , the partially or-
dered set of equivalence classes of normal series of submodules of M is directed:
that is, any two normal series admit equivalent refinements.
Proof. For a proof in a context which simultaneously generalizes that of mod-
ules and groups, see e.g. [J2, p. 106]. 
For an R-module M , a composition series is a maximal element in the partially
ordered set of normal series: that is, a composition series which admits no proper
refinement.
Exercise 8.11. Show: a normal series {Mi }ni=0 for an R-module M is a com-
position series iff for all 1 ≤ i ≤ n, the factor module Mi /Mi−1 is simple.
Theorem 8.13. (Jordan-Hölder) Let M be an R-module. Then any two com-
position series for M are equivalent: up to a permutation, their associated factor
series are term-by-term isomorphic.
Proof. This is an immediate consequence of Schreier Refinement: any two
normal series admit equivalent refinements, but no composition series admits a
proper refinement, so any two composition series must already be equivalent. 
Thus for a module M which admits a composition series, we may define the length
`(M ) of M to be the length of any composition series. One also speaks of the
Jordan-Hölder factors of M or the composition factors of M, i.e., the unique
multiset of isomorphism classes of simple R-modules which must appear as the suc-
cessive quotients of any composition series for M .
5. THE KRULL-SCHMIDT THEOREM 167

If a module does not admit a composition series, we say that it has infinite length.

And now a basic question: which R-modules admit a composition series?


Exercise 8.12. a) Show: any finite2 module admits a composition series.
b) Show that if a module M admits a composition series, it is finitely generated.
c) Show that a Z-module M admits a composition series iff it is finite.
d) Let k be a field. Show that a k-module admits a composition series iff it is finitely
generated (i.e., iff it is finite-dimensional).
Exercise 8.13. An R-module M admits a composition series iff there exists
L ∈ Z+ such that every normal series in M has length at most L.
Theorem 8.14. For an R-module M , the following are equivalent:
(i) M is both Noetherian and Artinian.
(ii) M admits a composition series.
Proof. Assume (i). Since M satisfies (DCC), there is a minimal nonzero
submodule, say M1 . If M1 is a maximal proper submodule, we have a composition
series. Otherwise among all proper R-submodules strictly containing M1 , by (DCC)
we can choose a minimal one M2 . We continue in this way: since M also satisfies
(ACC) the process must eventually terminate, yielding a composition series.
(ii) =⇒ (i): This follows easily from Exercise 8.13. 
Exercise 8.14. Exercise 8.13 makes use of Schreier Refinement. Give a proof
that (ii) =⇒ (i) in Theorem 8.14 which is independent of Schreier Refinement.
(Suggestion: try induction on the length of a composition series.)
Proposition 8.15. Let 0 → M 0 → M → M 00 → 0 be a short exact sequence of
R-modules. Then:
a) M admits a composition series iff both M 0 and M 00 admit composition series.
b) If M admits a composition series, then
`(M ) = `(M 0 ) + `(M 00 ).
Exercise 8.15. Prove Proposition 8.15.
Although it will not play a prominent role in our course, the length of an R-
module M is an extremely important invariant, especially in algebraic geometry:
it is is used, among other things, to keep track of intersection multiplicities and to
quantitatively measure the degree of singularity of a point.

5. The Krull-Schmidt Theorem


The material in this section follows [J2, §3.4] very closely. In particular, very ex-
ceptionally for us – but as in loc. cit. – in this section we work with left modules
over a possibly noncommutative ring R. The reason: not only does the desired
result carry over verbatim to the noncommutative case (this is not in itself a good
enough reason, as the same holds for a positive proportion of the results in these
notes) but the proof requires us to consider noncommutative rings!

2Recall that by a “finite module” we mean a module whose underlying set is finite!
168 8. NOETHERIAN RINGS

A module M is decomposable if there are nonzero submodules M1 , M2 ⊂ M


such that M = M1 ⊕ M2 ; otherwise M is indecomposable.
Theorem 8.16. (Krull-Schmidt) Let M be an R-module of finite length. Lm Then:
a) There are indecomposable submodules M1 , . . . , Mm such that M =L i=1 Mi .
n
b) If there are indecomposable submodules N1 , . . . , Nn such that M = i=1 Ni , then
m = n and there exists a bijection σ of {1, . . . , n} such that for all i, Mi ∼
= Nσ(i) .
The Proof of Theorem 8.16a) is easy, and we give it now. If M is a finite length
module and we write M = M1 ⊕ M2 then 0 < `(M1 ), `(M2 ) < `(M ). Thus an
evident induction argument shows that any sequence of moves, each one of which
splits a direct summand of M into two nontrivial direct subsummands of M , must
terminate after finitely many steps, leaving us with a decomposition of M into a
finite direct sum of indecomposable submodules. 

As one might suspect, the second part of Theorem 8.16 concerning the unique-
ness of the indecomposable decomposition is more subtle. Indeed, before giving the
proof we need some preparatory considerations on endomorphism rings of modules.
Proposition 8.17. For an R-module M , the following are equivalent:
(i) M is decomposable.
(ii) The (possibly noncommutative, even if R is commutative) ring EndR (M ) =
HomR (M, M ) has a nontrivial idempotent, i.e., an element e 6= 0, 1 with e2 = e.
Exercise 8.16. Prove Proposition 8.17.
A not-necessarily-commutative ring R is local if the set of nonunits R \ R× forms
a two-sided ideal of R.
Exercise 8.17. Let R be a local, not necessarily commutative ring.
a) Show that R 6= 0.
b) Show that R has no nontrivial idempotents.
An R-module M is strongly indecomposable if EndR (M ) is local. Thus it follows
from Proposition 8.17 and Exercise 8.17 that a strongly indecomposable module is
indecomposable.
Example 8.18. The Z-module Z is indecomposable: any two nonzero submod-
ules (a) and (b) have a nontrivial intersection (ab). On the other hand EndZ (Z) = Z
is not a local ring, so Z is not strongly indecomposable.
Thus “strongly indecomposable” is, in general, a stronger concept than merely “in-
decomposable”. Notice though that the Krull-Schmidt theorem applies only to
finite length modules – equivalently to modules which are both Noetherian and
Artinian – and Z is not an Artinian Z-module. In fact, it shall turn out that any
finite length indecomposable module is strongly indecomposable, and this will be a
major step towards the proof of the Krull-Schmidt Theorem.
But we are not quite ready to prove this either! First some Fitting theory.

For an R-module M and f ∈ EndR (M ), we put



\
f ∞ (M ) = f n (M ).
n=1
5. THE KRULL-SCHMIDT THEOREM 169

The set f ∞ (M ) is the intersection of a descending chain


M ⊃ f (M ) ⊃ f 2 (M ) ⊃ . . . ⊃ f n (M ) ⊃ . . .
of submodules of M , and is thus an f -stable submodule of M . The restriction of f
to f ∞ (M ) is surjective. Moreover, if M is an Artinian module, there exists s ∈ Z+
such that f s (M ) = f s+1 (M ) = . . ..
Exercise 8.18. Find a commutative ring R, an R-module M and f ∈ EndR (M )
such that for no n ∈ Z+ is the submodule f n (M ) f -stable.
Similarly, for M and f as above, we put

[
f−∞ (0) = ker f n .
n=1
Here each ker f n is an f -stable submodule of M on which f is nilpotent. The set
f−∞ (0) is the union of an ascending chain of submodules
0 ⊂ ker f ⊂ ker f 2 ⊂ . . . ⊂ ker f n ⊂ . . .
of M and is thus an f -stable submodule of M on which f acts as a nil endomor-
phism: i.e., every element of M is killed by some power of f . Moreover, if M is a
Noetherian module, there exists t ∈ Z+ such that ker f t = ker f t+1 = . . . and thus
f is a nilpotent endomorphism of f−∞ (0).
Exercise 8.19. Find a commutative ring R, an R-module M and f ∈ EndR (M )
such that f is not a nilpotent endomorphism of f−∞ (0).
Theorem 8.19. (Fitting’s Lemma) Let M be a finite length module over the
not necessarily commutative ring R, and let f ∈ EndR (M ).
a) There exists a Fitting Decomposition
(17) M = f ∞ (M ) ⊕ f −∞ (0).
b) f |f ∞ (M ) is an isomorphism and f |f−∞ (0) is nilpotent.
Proof. Since M has finite length it is both Noetherian and Artinian. Thus
there exists r ∈ Z+ such that
f r (M ) = f r+1 (M ) = . . . = f ∞ (M )
and
ker f r = ker f r+1 = . . . = f−∞ (0).
Let x ∈ f ∞ (M ) ∩ f−∞ (0). Then there is y ∈ M such that x = f r (y); moreover
0 = f r (x) = f 2r (y). But f 2r (y) = 0 implies x = f r (y) = 0, so f ∞ (M )∩f−∞ (0) = 0.
Let x ∈ M . Then f r (x) ∈ f r (M ) = f 2r (M ), so there exists y ∈ M with
f r (x) = f 2r (y) and thus f r (x − f r (y)) = 0. so
x = f r (y) + (x − f r (y)) ∈ f ∞ (M ) + f −∞ (0),
completing the proof of part a). As for part b), we saw above that the restriction of
f to f ∞ (M ) is surjective. It must also be injective since every element of the kernel
lies in f −∞ (0). Thus f |f ∞ (M ) is an isomorphism. Finally, as observed above, since
f−∞ (0) = ker f r , f |f−∞ (0) is nilpotent. 
Corollary 8.20. Let M be a finite length indecomposable R-module. Then ev-
ery f ∈ EndR (M ) is either an automorphism or nilpotent. Moreover M is strongly
indecomposable.
170 8. NOETHERIAN RINGS

Proof. Since M is indecomposable, Fitting’s Lemma implies that for f ∈


EndR (M ) we must have either M = f ∞ (M ) – in which case f is an automorphism
– or M = f −∞ (0) – in which case f is nilpotent. We must show that
I := EndR (M ) \ EndR (M )×
is a two-sided ideal of EndR (M ). We observe that I consists precisely of the nilpo-
tent elements of EndR (M ).
For f ∈ I and g ∈ EndR (M ), since f is neither injective nor surjective, we have
gf is not injective and f g is not surjective, hence f g, gf ∈ I.
Let f1 , f2 ∈ I, and seeking a contradiction we suppose that f1 + f2 ∈ / I, so
u := f1 + f2 ∈ EndR (M )× . For i = 1, 2, we put hi = fi u−1i , so h1 + h2 = 1. Then
h2 is non-invertible hence nilpotent: we have hn2 = 0 for some n ∈ Z+ and thus
(1 − h2 )(1 + h2 + . . . + hn−1
2 ) = 1 = (1 + h2 + . . . + hn−1
2 )(1 − h2 ),
contradicting that h1 = 1 − h2 ∈ I 
Lemma 8.21. Let M be a nonzero R-module and N an indecomposable R-
module. Suppose we have homomorphisms f : M → N, g : N → M such that gf
is an automorphism of M . Then both f and g are isomorphisms.
Proof. Let h = (gf )−1 , l = hg : N → M and e = f l : N → N . Then
lf = hgf = 1M and e2 = f lf l = f 1M l = f l = e. Since M is indecomposable, either
e = 1 or e = 0, and the latter implies 1M = 12M = lf lf = lef = 0, i.e., M = 0. So
f l = e = 1N , so f is an isomorphism and thus so too is (f (gf )−1 )−1 = g. 
8.22. Let M ∼
L m
Theorem
Ln = N be isomorphic modules, and let M = i=1 Mi and
0
N = i=1 Ni with each Mi strongly indecomposable and each Ni indecomposable.
Then m = n and there is a bijection σ of {1, . . . , m} such that for all i, Mi ∼
= Nσ(i) .
Proof. By induction on m: m = 1 is clear. Suppose the result holds for all
direct sums of fewer than m strongly indecomposable submodules.
Step 1: Let e1 , . . . , em ∈ EndR (M ) and f1 , . . . , fn ∈ EndR (N ) be the idempotent
elements corresponding to the given direct sum decompositions (i.e., projection

onto the corresponding factor). Let g : M → N , and put
hj := fj ge1 ∈ HomR (M, N ), kj = e1 g −1 fj ∈ HomR (N, M ), 1 ≤ j ≤ n.
Then
n
X X X
kj hj = e1 g −1 fj ge1 = e1 g −1 fj ge1 = e1 g −1 1N ge1 = e1 .
j=1 j j

The restrictions of e1 and kj hj to M1 stabilize M1 so may be regarded as endomor-


phisms of M1 , say e01 and (kj hj )0 , and we have
n
X
(kj hj )0 = e01 = 1M1 .
j=1

By assumption EndR M1 is local, so for at least one j, (kj hj )0 is a unit, i.e., an


automorphism of M1 . By reordering the Nj ’s we may assume that j = 1, so
(k1 h1 )0 ∈ AutR M1 . We may regard the restriction h01 of h1 to M1 as a homomor-
phism from M1 to N1 and similarly the restriction k10 of k1 to N1 as a homomorphism
from N1 to M1 , and then k10 h01 = (k1 h1 )0 is an automorphism. By Lemma 8.21,
6. SOME IMPORTANT TERMINOLOGY 171

∼ ∼
h01 = (f1 ge01 ) : M1 → N1 and k10 = (e1 g −1 f1 )0 : N1 → M1 .
Step 2: We claim that
M m
(18) M = g −1 (N1 ) ⊕ Mi .
i=2
−1
Lm −1
Lmthis, let x ∈ g N1 ∩ ( i=2 Mi ), so x = g y for some y ∈ N1 . Because
To see
x ∈ i=2 Mi , e1 x = 0. Thus
0 = e1 x = e1 g −1 y = e1 g −1 f1 y = k1 y = k10 y.
Since k10 is an isomorphism, Lm y = 0 and thus x = 0, so the sum in (18) is direct.
Now put M 0 = g −1 (N1 ) ⊕ i=2 Mi , so we wish to show M 0 = M . Let x ∈ g −1 N1 .
Then x, e2 x, . . . , em x ∈ M 0 , so e1 x = (1 − e2 − . . . − em )x ∈ M 0 . So
M 0 ⊃ e1 g −1 N1 = e1 g −1 f1 N1 = k1 N1 = k10 N1 = M1
Lm
and thus M 0 ⊃ i=1 Mi = M .

Step 3: The isomorphism g : M → N carries g −1 N1 onto N1 hence induces an
M sim
isomorphism g−1 N1 → N/N1 . Using Step 2, we have
n m
M N ∼ M ∼M
Ni = = −1 = Mi .
j=2
N1 g N1 i=2

We are done by induction. 


Exercise 8.20. Please confirm that we have proved the Krull-Schmidt Theo-
rem!
Exercise 8.21. Let M and N be R-modules such that M × M ∼
= N × N.
a) If M and N are both of finite length, show that M ∼
= N.
b) Must we have M ∼= N in general?
Part b) is far from easy! If you give up, see [Co64].

6. Some important terminology


All we aspire to do in this section is to introduce some terminology, but it is so
important that we have isolated it for future reference.

Let R be a ring and p a prime ideal of R. The height of p is the supremum


of all lengths of finite chains of prime ideals of the form p0 ( p1 ( . . . ( pn = p
(the length of the indicated chain being n; i.e., it is the number of (’s appear-
ing, which is one less than the number of elements). Thus the height is either a
non-negative integer or ∞; the latter transpires iff there exist arbitrarily long finite
chains of prime ideals descending from p (and of course, this need not imply the
existence of an infinite chain of prime ideals descending from p).

A prime ideal of height 0 is called a minimal prime. In a domain R, the unique


minimal prime is (0), so the concept is of interest only for rings which are not
domains. If I is a proper ideal of R, we also speak of a minimal prime over I,
which means a prime p ⊃ I such that there is no prime ideal q with I ( q ( p. Note
that p is a minimal prime over I iff p is a minimal prime in the quotient ring R/I.
This remark simultaneously explains the terminology “minimal over” and gives a
172 8. NOETHERIAN RINGS

hint why it is useful to study minimal prime ideals even if one is ultimately most
interested in domains.

The dimension of a ring R is the supremum of all the heights of its prime ideals.
The full proper name here is Krull dimension of R, which is of course useful
when one has other notions of dimension at hand. Such things certainly do exist
but will not be considered here. Moreover, as will shortly become apparent, the
need to include Krull’s name here so as to ensure that he gets proper recognition
for his seminal work in this area is less than pressing. Therefore we use the full
name “Krull dimension” only rarely as a sort of rhetorical flourish.

One also often speaks of the codimension of a prime ideal p of R, which is the
dimension of R minus the height of p. This is especially natural in applications
to algebraic geometry, of which the present notes allude to only in passing. Note
that this is not necessarily equal to the Krull dimension of R/p – or what is the
same as that, the maximal length of a finite chain of prime ideals ascending from p
– although in reasonable applications, and especially in geometry, one is certainly
entitled to hope (and often, to prove) that this is the case.

Remark: All of these definitions would make perfect sense for arbitrary partially
ordered sets and their elements, but the terminology is not completely consistent
with order theory. Namely, the height of an element in an arbitrary partially or-
dered set is defined as the supremum of lengths of chains descending from that
element, but the order theorists would cringe to hear the supremum of all heights
of elements called the “dimension” of the partially ordered set. They would call
that quantity the height of the partially ordered set, and would reserve dimension
for any of several more interesting invariants. (Roughly, the idea is that a chain
of any finite length is one-dimensional, whereas a product of d chains should have
dimension d.)

7. Introducing Noetherian rings


The following is probably the most important single definition in all of ring theory.

A ring R is said to be Noetherian if the partially ordered set I(R) of all ideals of
R satisfies the ascending chain condition.
Exercise 8.22. Insert exercise here.
Theorem 8.23. A finitely generated module over a Noetherian ring is Noether-
ian.
Proof. If M is a finitely generated module over R, then we may represent it
as Rn /K for some submodule K of Rn . An immediate corollary of the preceding
theorem is that finite direct sums of Noetherian modules are Noetherian, and by
assumption R itself is a Noetherian R-module, hence so is Rn and hence so is the
quotient Rn /K = M . 
Thus so long as we restrict to Noetherian rings, submodules of finitely generated
modules remain finitely generated. This is extremely useful even in the case of
R = Z: a subgroup of a finitely generated commutative group remains finitely
8. THEOREMS OF EAKIN-NAGATA, FORMANEK AND JOTHILINGAM 173

generated. Needless(?) to say, this does not hold for all noncommutative groups,
e.g. not for a finitely generated free group of rank greater than 1.
Theorem 8.24. (Characterization of Noetherian rings) For a ring R, the
following are equivalent:
(i) Every nonempty set of ideals of R has a maximal element.
(ii) There are no infinite ascending chains
I1 ( I2 ( . . . ( In ( . . .
of ideals of R.
(iii) Every ideal of R is finitely generated.
(iv) Every prime ideal of R is finitely generated.
Proof. (i) ⇐⇒ (ii) is a special case of Proposition 8.1.
(ii) ⇐⇒ (iii) is a special case of Exercise 8.2.
(iii) ⇐⇒ (iv) is Cohen’s Theorem (Theorem 4.30). 
Exercise 8.23. Suppose a ring R satisfies the ascending chain condition on
prime ideals. Must R be Noetherian?
Proposition 8.25. Let R be a Noetherian ring.
a) If I is any ideal of R, the quotient R/I is Noetherian.
b) If S ⊂ R is any multiplicative subset, the localization S −1 R is Noetherian.
Proof. Any ideal of R/I is of the form J/I for some ideal J ⊃ I of R.
By assumption J is finitely generated, hence J/I is finitely generated, so R/I is
Noetherian. A similar argument holds for the localization; details are left to the
reader. 
Q
Exercise 8.24. Let k be a field, let S be an infinite set, and put R = s∈S k,
i.e., the infinite product of #S copies of k. Show that R is not Noetherian, but the
localization Rp at each prime ideal is Noetherian.
Thus Noetherianity is a localizable property but not a local property.

8. Theorems of Eakin-Nagata, Formanek and Jothilingam


In 1968, P.M. Eakin, Jr. [Ea68] and M. Nagata [Na68] independently showed that
if a ring R admits an extension ring S which is Noetherian and finitely generated
as an R-module, then R is Noetherian.
Several years later, E. Formanek [Fo73] gave a stronger result. His improve-
ment is a nice instance of the philosophy of “modulization”: where possible one
should replace theorems about rings with theorems about modules over rings. He
writes: “The object of this paper is to present a simple and elementary proof of
the Eakin-Nagata theorem which generalizes the original version in a new direction.
The proof is essentially a contraction of Eakin’s proof as presented by Kaplansky in
[K, Exc. 14-15, p. 54] based on the observation that much of the proof disappears
if one is not ‘handicapped’ by the hypothesis that T is a ring.”
More recently, P. Jothilingam [Jo00] gave a result which simultaneously gen-
eralizes Formanek’s Theorem and Cohen’s Theorem that a ring in which all prime
ideals are finitely generated is Noetherian. Finally(?), several years ago A. Naghipour
[Na05] found a significantly shorter, simpler proof of Jothilingam’s Theorem, which
we will present here. All in all, this provides a nice case study of how even very
174 8. NOETHERIAN RINGS

basic results get improved and simplified as time passes.


Having told the story in correct chronological order, we now reverse it: we will
prove Jothilingam’s Theorem and swiftly deduce the earlier results as corollaries.
First a couple of easy preliminaries.
Lemma 8.26 (Kaplansky). For a ring R, the following are equivalent:
(i) R is Noetherian.
(ii) R admits a faithful Noetherian module.
Proof. (i) =⇒ (ii): If R is Noetherian, then R is a faithful Noetherian R-
module.
(ii) =⇒ (i): Let M be a faithful Noetherian R-module. In particular M is finitely
generated, say by x1 , . . . , xn . Let ϕ : R → M n by r 7→ (rx1 , . . . , rxn ). Since
M is Noetherian, so is M n , and since M is faithful, ϕ is injective, and thus R is
isomorphic to a submodule of a Noetherian module, hence Noetherian. 
Exercise 8.25. a) Show: any ring R admits a Noetherian module.
b) Show: if M is a Noetherian R-module, R/ ann M is Noetherian.
Let M be an R-module. An R-submodule of M is extended if it is of the form
IM for some ideal I of R. This is a generalization of a previous use of the term:
if ι : R → T is a map of rings, then the extended ideals of T are those of the form
ι∗ I = IT for an ideal I of R.
Proposition 8.27. For a finitely generated R-module M , let EM be the family
of extended submodules of M , partially ordered under inclusion. the following are
equivalent:
(i) EM is Noetherian: i.e., extended submodules satisfy (ACC).
(ii) Every extended submodule of M is finitely generated.
Proof. ¬ (ii) =⇒ ¬ (i): Let I be an ideal of R such that IM is not finitely
generated. Let a1 ∈ I. Then, since M is finitely generated, a1 M is a finitely
generated submodule of IM , hence proper: there exists a2 ∈ I such that a1 M (
ha1 , a2 iM . Again, ha1 , a2 iM is finitely generated, so is proper in IM . Continuing
in this way we get a sequence {an }∞ n=1 in I such that
a1 M ( ha1 , a2 iM ( . . . ( ha1 , . . . , an iM ( . . . ,
so EM is not Noetherian.
(ii) =⇒ P (i): Let I1 M ⊆ I2PM ⊆ . . . ⊆ In M ⊆ . . . be an ascending chain in EM .
Let N = n In M and I = n In , so N = IM ∈ EM . By assumption, N is finite
generated, so there is n ∈ Z+ with N = I1 M + . . . + In M . Since Ik M ⊂ Ik+1 M for
all k, N = In M and thus In M = In+k M for k ∈ N: the chain stabilizes at n. 
Theorem 8.28 (Jothilingam). For a finitely generated R-module M , the fol-
lowing are equivalent:
(i) M is Noetherian.
(ii) For every prime ideal p of R, the submodule pM is finitely generated.
Proof. We follow [Na05].
(i) =⇒ (ii): If M is Noetherian, then every submodule of M is finitely generated.
¬ (i) =⇒ ¬ (ii): Suppose M is not Noetherian: we will find a prime ideal p of R
such that pM is infinitely generated.
Step 0: Since the union of a chain of infinitely generated submodules of M is
8. THEOREMS OF EAKIN-NAGATA, FORMANEK AND JOTHILINGAM 175

an infinitely generated submodule of M , by Zorn’s Lemma there is a submodule


N ⊂ M maximal with respect to being infinitely generated.
Step 1: Let p = ann(M/N ) = {x ∈ R | xM ⊂ N }. We will show that p is a
prime ideal: indeed, seeking a contradiction suppose there are a, b ∈ R \ p such
that ab ∈ p. Then N + aM, B + bM ) N so are both finitely generated: write
N + aM = hn1 + am1 , . . . , n` + am` i with ni ∈ N , mi ∈ M . Put
L = {m ∈ M : am ∈ N };
then L is an R-submodule of M containing N and bM and hence also N +bM ) N ,
so L is finitely generated. We claim
`
X
N= Rni + aL.
i=1

If so, then N is finitely generated, a contradiction, and thus p is prime. Since


P`
abM ⊂ N , we have i=1 Rni + aL ⊂ N . Conversely, let y ∈ N . Since y ∈ N + aM ,
there are b1 , . . . , b` ∈ R such that
`
X `
X `
X
y= bi (ni + ami ) = bi n i + a bi mi .
i=1 i=1 i=1

Thus
`
X `
X
a bi mi = y − bi ni ∈ N,
i=1 i=1
P` P`
so i=1 bi mi ∈ L and y ∈ i=1 Rni + aL.
Step 2: For x ∈ M , write x for the canonical image of x in M/N . Now we use
that M Tis finitely generated: write M = hx1 , . . . , xn iR , so M/N = hx1 , . . . , xn iR ,
n
so p = i=1 ann Rxi . Because p is prime, we must have p = ann Rxj for some j.
Since N + Rxi ) N , N + Rxi is finitely generated, say by y1 + r1 xj , . . . , yk + rk xj ,
with yi ∈ N , ri ∈ R. Arguing as in Step 1 we get
k
X
N= Ryi + pxj .
i=1

Since pM ⊂ N , we have
k
X k
X k
X
N= Ryi + pxj ⊂ Ryi + pM ⊂ Ryi + N ⊂ N,
i=1 i=1 i=1

and thus
k
X
(19) N= Ryi + pM.
i=1

Since N is infinitely generated, (19) implies pM is infinitely generated. 

Corollary 8.29. (Formanek’s Theorem) Let R be a ring, and let M =


ha1 , . . . , an i be a faithful finitely generated R-module. Suppose M satisfies (ACC)
on “extended submodules” – i.e., submodules of the form IM for I an ideal of R.
Then M is Noetherian, hence so is R.
176 8. NOETHERIAN RINGS

Proof. By Proposition 8.27, all extended submodules are finitely generated,


hence a fortiori all submodules of the form pM for p ∈ Spec R are finitely generated.
By Theorem 8.28, M is Noetherian, and then by Lemma 8.26, R is Noetherian. 
Corollary 8.30. (Eakin-Nagata Theorem) Let R ⊂ S be an ring extension,
with S finitely generated as an R-module. Then R is Noetherian iff S is Noetherian.
Proof. =⇒ If R is Noetherian, then S is a finitely generated module over
a Noetherian ring so S is a Noetherian R-module. That is, (ACC) holds on R-
submodules of S, hence a fortiori it holds on S-submodules of S.
⇐ Apply Formanek’s Theorem with M = S. 
Exercise 8.26. Investigate the possibility of proving Jothilingam’s Theorem
using the Prime Ideal Principle of §4.5.

9. The Bass-Papp Theorem


We now present a beautiful characterization of Noetherian rings in terms of proper-
ties of injective modules, due independently to Z. Papp [Pa59] and H. Bass [Ba59].
Theorem 8.31. (Bass-Papp) For a ring R, the following are equivalent:
(i) A direct limit of injective modules is injective.
(ii) A direct sum of injective modules is injective.
(iii) A countable direct sum of injective modules is injective.
(iv) R is Noetherian.
Proof.
(i) =⇒ (ii): A direct sum is a kind of direct limit.
(ii) =⇒ (iii) is immediate.
(iii) =⇒ (iv): LetSI1 ⊂ I2 ⊂ . . . ⊂ In ⊂ . . . be an infinite ascending chain of ideals
of R, and let I = n In . We define

M
E= E(R/In ).
n=1

For n ∈ Z+ , let fn : I → E(R/In ) beQthe composite Q∞ map I → R → R/In →


E(R/In ). There is then a unique map f : I → n=1 E(R/In ). But indeed, for
fixed x ∈ I, x lies in In for sufficiently large n and thus fn (x) = 0. It follows
each Q
that f actually lands in the direct sum, and we have thus defined a map
f : I → E.
By hypothesis, E is a countable direct sum of injective modules and therefore
injective, so f extends to an R-module map with domain all of R and is thus of the
form f (x) = xf (1) = xe for some fixed e ∈ E. Let N be sufficiently large so that
for n ≥ N , the nth component en of e is zero. Then for all x ∈ I,
0 = xen = fn (x) = x + In ∈ R/In ,
and thus x ∈ In . That is, for all n ≥ N , In = I.
(iv) =⇒ (i): let {Eα } be a directed system of injective modules with direct limit
E. For α ≤ β we denote the transition map from Eα to Eβ by ιαβ and the natural
map from Eα to E by ια . We will show E is injective by Baer’s Criterion (Theorem
3.20), so let I be any ideal of R and consider an R-module map f : I → E. Since
R is Noetherian, I is finitely generated, and it follows that there exists an index
10. ARTINIAN RINGS: STRUCTURE THEORY 177

α such that f (I) ⊂ ια (Eα ). Let M be a finitely generated submodule of Eα such


that f (I) ⊂ ια (M ). Consider the short exact sequence
α ι
0→K→M → f (I) → 0.
Since M is finitely generated and R is Noetherian, K is finitely generated. Moreover
K maps to 0 in the direct limit, so there exists β ≥ α such that ιαβ K = 0. Let
M 0 = ιαβ M , so by construction

ιβ : M 0 → f (I).
Taking g = ιβ |−1
M 0 ◦ f we get a map g : I → Eβ such that f = ιβ ◦ g. Since Eβ is
injective, g extends to a map G : R → Eβ and thus F = ιβ ◦ G extends f to R. 

10. Artinian rings: structure theory


A ring R which satisfies the descending chain condition (DCC) on ideals is called
Artinian (or sometimes, “an Artin ring”).
Exercise 8.27. a) Show: a ring with only finitely many ideals is Artinian.
b) Show: the ring of integers Z is not Artinian.
c) Show: a quotient of an Artinian ring is Artinian.
d) Show: a localization of an Artinian ring is Artinian.
Obviously any finite ring has only finitely many ideals and is Artinian. It is not
difficult to give examples of infinite rings with finitely many ideals. For instance,
let k be a field and let 0 6= f ∈ k[t]. Then R = k[t]/(f ) has only finitely many
ideals. Indeed, if we factor f = f1a1 · · · frar into irreducible factors, then the Chinese
Remainder Theorem gives
k[t]/(f ) ∼
= k[t]/(f a1 ) × . . . × k[t]/(f ar ).
1 r
Each factor ring k[t]/(fiai ) is a local ring with maximal ideal (f1 ), and the ideals
are precisely
(0) ( (fi )ai −1 ( . . . ( fi .
Since every
Qrideal in a product is a direct sum of ideals of the factors, there are then
precisely i=1 (ai + 1) ideals of R.

A bit of reflection reveals that – notwithstanding their very similar definitions –


requiring (DCC) on ideals of a ring is considerably more restrictive than the (ACC)
condition. For instance:
Proposition 8.32. A domain R is Artinian iff it is a field.
Proof. Obviously a field satisfies (DCC) on ideals. Conversely, if R is a
domain and not a field, there exists a nonzero nonunit element a, and then we have
(a) ) (a2 ) ) (a3 ) ) . . .. Indeed, if (ak ) = (al ), suppose k ≤ l and write l = k + n,
and then we have uak = ak an for some u ∈ A× , and then by cancellation we get
an = u, so an is unit and thus a is a unit, contradiction. 
The result collects several simple but important properties of Artinian rings.
Theorem 8.33. Let R be an Artinian ring.
a) R has dimension zero: prime ideals are maximal.
b) Therefore the Jacobson radical of R coincides with its nilradical.
c) R has only finitely many maximal ideals, say m1 , . . . , mn .
178 8. NOETHERIAN RINGS

Tn
d) Let N = i=1 mi be the nilradical. Then it is a nilpotent ideal: there exists
k ∈ Z+ such that N k = 0.
Proof. a) If p is a prime ideal of A, then A/p is an Artinian domain, which
by Proposition 8.32 is a field, so p is maximal.
b) By definition, the Jacobson radical is the intersection of all maximal ideals and
the nilradical is the intersection of all prime ideals. Thus the result is immediate
from part a).
c) Suppose mi is an infinite sequence of maximal ideals. Then
R ) m1 ) m1 ∩ m2 · · ·
is an infinite descending chain. Indeed, equality at any step would mean mN +1 ⊃
TN QN
i=1 mi = i=1 mi , and then since mN +1 is prime it contains mi for some 1 ≤ i ≤
N , contradiction.
d) By DCC, it must be the case that there exists some k with N k = N k+n for all
n ∈ Z+ . Put I = N k . Suppose I 6= 0, and let Σ be the set of ideals J such that
IJ 6= 0. Evidently Σ 6= ∅, for I ∈ Σ. By DCC we are entitled to a minimal element
J of Σ. There exists 0 6= x ∈ J such that xI 6= 0. For such an x, we have (x) ∈ Σ
and by minimality we must have J = (x). But (xI)I = xI 6= 0, so xI ⊂ (x) and
thus xI = (x) by minimality. So there exists y ∈ I with xy = x and thus we have
(20) x = xy = xy 2 = . . . = xy k = . . . .
But y ∈ I ⊂ N , so y is nilpotent and (20) gives x = 0, a contradiction. 
Lemma 8.34. Suppose that inQa ring R there exists a finite sequence m1 , . . . , mn
of maximal ideals such that 0 = i mi . Then R is Noetherian iff it is Artinian.
Proof. Recall from §8.4 that an R-module M is Noetherian and Artinian iff
it has finite length. Consider
0 = m1 · · · mn ( m1 · · · mn−1 ( . . . ( m1 ( R.
For R to have finite length, it is necessary and sufficient that each quotient
Qi = m1 · · · mi−1 /m1 · · · mi
be a finite length R-module. (Since we do not assume that the mi ’s are distinct, it is
possible – and harmless – that some Qi ’s may be zero.) But each Qi is canonically
a module over the field R/mi , i.e., a vector space, so it has finite length iff it is
finite-dimensional. So we win: R is Noetherian iff each Qi is iff each Qi is finite-
dimensional iff each Qi is Artinian iff R is Artinian. 
Theorem 8.35. (Akizuki-Hopkins) For a ring R, the following are equivalent:
(i) R is Artinian.
(ii) R is Noetherian, and prime ideals are maximal.
Proof. (i) =⇒ (ii): Suppose R is Artinian. By Theorem 8.33, prime ideals
in R are maximal, so it suffices to show that R is Noetherian. Let m1 ,Q . . . , mn be
n
the distinct maximal ideals of R. For any k ∈ Z+ we have (using CRT) i=1 mki =
Tn k
( i=1Qmi ) . Applying Theorem 8.33d), this shows that for sufficiently large k we
n
have i=1 mki = 0. We can now apply Lemma 8.34 to conclude that R is Artinian.
(ii) =⇒ (i): Suppose R is Noetherian and zero-dimensional. A bit later on
(sorry!) we will see that any Noetherian ring has only finitely many minimal prime
ideals (Theorem 10.13 and again in Corollary 13.19), so R has only finitely many
10. ARTINIAN RINGS: STRUCTURE THEORY 179

minimal
Tnprime ideals, each of which is maximal by zero-dimensionality. Therefore
N = i=1 mi is the nilradical of a Noetherian
Qn ring, hence a nilpotent ideal by
Proposition 4.11. As above, we deduce that i=1 mki = 0 for sufficiently large k.
By Lemma 8.34, R is Artinian. 

Exercise 8.28. Consider the ring R = C[x, y]/(x2 , xy, y 2 ) = C[x, y]/I.
a) Show that dimC R = 3 and that a C-basis is given by 1 + I, x + I, y + I.
b) Deduce that R is Artinian.
c) Show that the proper ideals of R are precisely the C-subspaces of hx + I, y + IiC .
d) Deduce that R has infinitely many ideals.
Exercise 8.29. Let k be a field and A = k[{xi }∞ i=1 ] a polynomial ring over k
in a countable infinite number of indeterminates. Let m = ({xi }) be the ideal of all
polynomials with zero constant term, and put R = A/m2 . Show that R is a ring
with a unique prime ideal which is not Noetherian (so also not Artinian).
Exercise 8.30. Let n ∈ Z+ . Suppose R is a Noetherian domain with exactly
n prime ideals. Must R be Artinian?
Proposition 8.36. Let (R, m) be a Noetherian local ring.
a) Either:
(i) mk 6= mk+1 for all k ∈ Z+ , or
(ii) mk = 0 for some k.
b) Moreover, condition (ii) holds iff R is Artinian.
Proof. a) Suppose there exists k such that mk = mk+1 . By Nakayama’s
Lemma, we have mk = 0. If p is any prime ideal of R, then mk ⊂ p, and taking
radicals we have m ⊂ p, so p = m and R is a Noetherian ring with a unique prime
ideal, hence an Artinian local ring. b) If R is Artinian, then (i) cannot hold, so (ii)
must hold. Conversely, if (ii) holds then m is a nil ideal, hence contained in the
intersection of all prime ideals of R, which implies that m is the only prime ideal
of R, and R is Artinian by the Akizuki-Hopkins theorem. 

Theorem 8.37. Let R be an Artinian ring.


a) There are n ∈ Z+ and local Artinian rings R1 , . . . , Rn such that R ∼
Qn
=Q i=1 Ri .
b) Moreover, the decomposition is unique in the sense that if R ∼
m
= j=1 Sj is
another decomposition, then n = m and there exists a permutation σ of {1, . . . , n}
such that Ri ∼
= Sσ(i) for all i.
Proof. a) Let (mi )ni=1 beQ
the distinct maximal ideals of R. We have seen that
n
there exists k ∈ Z+ such that
T ki=1 Q mki = 0. By Proposition 4.18, the ideals mki are
pairwise comaximal, so so i mi = i mki . Therefore by CRT the natural mapping
n
Y R
R→ k
i=1
mi

is an isomorphism. Each mRk is local Artinian, so this gives part a).


i
b) The proof requires primary decomposition, so must be deferred to §10.5. 

Exercise 8.31. Let R be an Artinian ring.


a) Show: every element of R is either a unit or a zero divisor.
b) Show: R is its own total fraction ring.
180 8. NOETHERIAN RINGS

Exercise 8.32. Let R be a ring, and let J(R) be its Jacobson radical. The
following are equivalent:
(i) R is semilocal, i.e., MaxSpec R is finite.
(ii) The ring R/J(R) is a finite product of fields.
(iii) The ring R/J(R) has only finitely many ideals.
(iv) The ring R/J(R) is Artinian.

11. The Hilbert Basis Theorem


The following result shows in one fell swoop that the majority of the rings that one
encounters in classical algebraic geometry and number theory are Noetherian.
Theorem 8.38. (Hilbert Basis Theorem) If R is Noetherian, so is R[t].
Proof. Seeking a contradiction, suppose J is an ideal of R[t] which is not
finitely generated. We inductively construct a sequence f0 , f1 , . . . , fn , . . . of ele-
ments of J and a sequence of ideals Jn = hf0 , . . . , fn i of R[t] as follows: f0 = 0,
and for all i ∈ N, fi+1 is an element of minimal degree in J \ Ji . Thus for all
positive integers i we have deg fi ≤ deg fi+1 . Moreover, for all i ∈ Z+ let ai be the
leading coefficient of fi , and let I be the ideal ha1 , a2 , . . . , aN , . . .i of R. However,
R is Noetherian, so there exists N ∈ Z+ such that I = ha1 , . . . , aN i. In particular,
there are u1 , . . . , uN ∈ R such that aN +1 = u1 a1 + . . . + uN aN . Define
N
X
g= ui fi tdeg fN +1 −deg fi .
i=1
Since g ∈ JN and fN +1 ∈ J \ JN , we have fN +1 − g ∈ J \ JN . Moreover g and fN +1
have the same degree and the same leading term, so deg fN +1 −g < deg fN +1 , hence
fN +1 does not have minimal degree among polynomials in J \JN , contradiction. 
Exercise 8.33. Prove the converse of the Hilbert Basis Theorem: if R is a
ring such that either R[t] or R[[t]] is Noetherian, then R is Noetherian.
Corollary 8.39. A finitely generated algebra over a Noetherian ring is Noe-
therian.
Proof. Let R be Noetherian and S a finitely generated R-algebra, S ∼ = R[t1 , . . . , tn ]/I
for some n ∈ Z+ and some ideal I. By induction on the Hilbert Basis Theorem,
the ring R[t1 , . . . , tn ] is Noetherian, hence so is its quotient ring S. 
Theorem 8.40. Let R be a ring, let P be a prime ideal of R[[t]], and let p be
the set of constant coefficients of elements of P.
a) Suppose that for some k ∈ N, p can be generated by k elements. Then P can be
generated by k + 1 elements. Moreover, if t ∈/ P, P can be generated by k elements.
b) If R is Noetherian, then so is R[[t]].
Proof. Let q : R[[t]] → R[[t]]/(t) = R be the quotient map, so p = q∗ P.
a) Suppose p = ha1 , . . . , ak i, and let I be the ideal ha1 , . . . , ak , ti of R[[t]].
∈ P, we claim I = P, which suffices. That I ⊂ P is clear; conversely,
Case 1: If t P

writing f = n=0 an tn ∈ P as a0 + t(a1 + a2 t + . . .) shows f ∈ I.
Case 2: Suppose t ∈ / P. Let f1 , . . . , fk ∈ P with constant terms P∞ a1 , . . . , ak , respec-
tively. We claim P = hf1 , . . . , fk i. To see this, let g1 = n=0 bn tn ∈ P. Since
b0 ∈ p, there are r1,1 , . . . , rk,1 ∈ R with
b0 = r1,1 a1 + . . . + rk,1 ak ,
12. THE KRULL INTERSECTION THEOREM 181

and thus
g1 − (r1,1 f1 + . . . + rk,1 fk ) = tg2
for some g2 ∈ R[[t]]. Since P is prime, tg2 ∈ P and t ∈ / P, we must have g2 ∈ P.
Applying the above argument to g2 we find r1,2 , . . . , rk,2 ∈ R and g3 ∈ P such that
g2 − (r1,2 f1 + . . . + r1,k
Pf∞k ) = tg3 . Continuing in this way, we generate, for 1 ≤ i ≤ k,
a power series hi = n=0 ri,n tn , such that
g = h1 f1 + . . . + hk fk ,
establishing the claim.
b) If R is Noetherian, then by part a) every prime ideal of R[[t]] is finitely generated.
By Cohen’s Theorem (Theorem 4.30), R[[t]] is Noetherian. 
Exercise 8.34. Show: for a ring R, the following are equivalent:
(i) R is Noetherian.
(ii) For all n ≥ 1, R[t1 , . . . , tn ] is Noetherian.
(iii) For all n ≥ 1, R[[t1 , . . . , tn ]] is Noetherian.
Exercise 8.35. Let k be a field, and consider the subring R = k[y, xy, x2 y, . . .]
of k[x, y]. Show that R is not Noetherian.
Therefore, a subring of a Noetherian ring need not be Noetherian. Thinking that
this ought to be the case is one of the classic “rookie mistakes” in commutative
algebra. In general though, it is the exception rather than the rule that a nice
property of a ring R is inherited by all subrings of R, and one gets used to this.

12. The Krull Intersection Theorem


12.1. Preliminaries on Graded Rings.

In the proof of the theorem of this section we will need a little fact about homoge-
neous polynomials. So here we discuss some rudiments of this theory by embedding
it into its natural context: graded rings. The notion of graded ring is of the utmost
importance in various applications of algebra, from algebraic geometry to algebraic
topology and beyond. It would certainly be nice to give a comprehensive exposition
of graded algebra but at the moment this is beyond the ambition of these notes, so
we content ourselves with the bare minimum needed for our work in the next section.

Let R be a ring, n ∈ Z+ , and denote by R[t] = R[t1 , . . . , tn ] the polynomial ring


in n indeterminates over R. For a polynomial P = P (t) in several variables, we
have the notion of the degree of P with respect to the variable ti : thinking of P
as an element of R[t1 , . . . , ti−1 , ti+1 , . . . , tn ][ti ] it is just the largest m such that the
i1
coefficient of tm
i is nonzero, as usual. For any monomial term cI t1 · · · tn we define
in

the total degree to be d = i1 + . . . + in .

A nonzero polynomial P = I cI ti11 · · · tinn is homogeneous if all of its monomial


P
terms have the same total degree, and this common number is called the degree
of the homogeneous polynomial P . By convention the zero polynomial is re-
garded as being homogeneous total degree d for all d ∈ N.

A general polynomial
P∞ P ∈ R[t] can be written as a sum of homogeneous poly-
nomials P = d=0 P d (t) with each Pd homogeneous of degree d (and of course
182 8. NOETHERIAN RINGS

Pd = 0 for all sufficiently large d). This sum is unique. One way to see this is
to establish the following more structural fact: for any d ∈ N, let P [t]d be the
set of all polynomials which are homogeneous of degree d. Then each P [t]d is an
R-submodule of P [t] and we have a direct sum decomposition

M
(21) P [t] = P [t]d .
d=0
Moreover, for all d1 , d2 ∈ N we have

(22) P [t]d1 · P [t]d2 ⊂ P [t]d1 +d2 .


In general, if R is aLring and S is an algebra admitting an R-module direct sum

decomposition S = d=0 Sd satisfying Sd1 · Sd2 ⊂ Sd1 +d2 , then we say that S is an
(N)-graded R-algebra. Taking R = Z we get the notion of a graded ring.
L∞
Exercise 8.36. Let S = d=0 Sd be a graded R-algebra. Show: the R-
submodule S0 is in fact an R-algebra.
L∞
Let S = d=0 Sd be a graded ring. We say that x ∈ S is homogeneous of degree
d if x ∈ Sd . An ideal I of S is homogeneous if it has a generating set I = hxi i
with each xi a homogeneous element.
Exercise 8.37. Let S be a graded R-algebra and let I be a homogeneous ideal
of S. Show:
M∞
S/I = (Sd + I)/I
d=0
and thus S/I is a graded R-algebra.
Now back to the case of polynomial rings.
Lemma 8.41. Let S be a graded ring, let f1 , . . . , fn be homogeneous elements
of S, and put I = hf1 , . . . , fn i. Let f ∈ I be homogeneous. Then there are homoge-
neous elements g1 , . . . , gn ∈ R such that
Xn
f= gi fi
i=1
and for all 1 ≤ i ≤ n,
deg gi = deg f − deg fi .
Proof. Since f ∈ I, there exist X1 , . . . , Xn ∈ S such that
f = X1 f1 + . . . + Xn fn .
P
For each 1 ≤ i ≤ n, let Xi = j xi,j with deg xi,j = j be the canonical decomposi-
tion of Xi into a sum of homogeneous elements: i.e., deg xi,j = j. Then
∞ X
X n
(23) f= xi,d−deg fi fi .
d=0 i=1
Since f is homogeneous of degree deg(f ), only the d = deg(f ) in the right hand
side of (23) is nonzero, so
Xn
f= xi,deg f −deg fi fi . 
i=1
12. THE KRULL INTERSECTION THEOREM 183

12.2. The Krull Intersection Theorem.


Theorem 8.42. Let R be a Noetherian T∞ ring, and I an ideal of R. Suppose
there is an element x of R such that x ∈ n=1 I n . Then x ∈ xI.

Proof. The following miraculously short proof is due to H. Perdry [Pe04].


Suppose I = ha1 , . . . , ar i. For each n ≥ 1, since x ∈ I n there is a homogeneous
degree n polynomial Pn (t1 , . . . , tr ) ∈ R[t1 , . . . , tr ] such that
x = Pn (a1 , . . . , an ).
By the Hilbert Basis Theorem (Theorem 8.38), the ring R[t1 , . . . , tr ] is Noetherian.
Therefore, definining Jn = hP1 , . . . , Pn i, there exists N such that JN = JN +1 . By
Lemma 8.41 we may write
PN +1 = QN P1 + . . . + Q1 PN ,
with Qi homogeneous of degree i > 0. Plugging in ti = ai for 1 ≤ i ≤ n, we get
x = PN +1 (a1 , . . . , an ) = x (Q1 (a1 , . . . , an ) + . . . + QN (a1 , . . . , aN )) .
Since each Qi is homogeneous of positive degree, we have Qi (a1 , . . . , an ) ∈ I. 

Corollary 8.43. Let I be an ideal in a Noetherian ring R. Suppose either


(i) R is a domain and I is a proper ideal; or
(ii) I T
is contained in the Jacobson radical J(R) of R.

Then n=1 I n = 0.
T∞
Proof. Either way, let x ∈ n=1 I n and apply Theorem 8.42 to obtain an
element a ∈ I such that x = xa. Thus (a − 1)x = 0. Under assumption (i), we
obtain either a = 1 – so I = R, contradicting the properness of I – or x = 0. Under
assumption (ii), a ∈ J(R) implies a − 1 ∈ R× , so that we may multiply through by
(a − 1)−1 , again getting x = 0.F 

T∞ n 8.38. (Suárez-Alvarez) Exhibit an ideal I in a Noetherian ring such


Exercise
that n=1 I ) {0}. (Hint: idempotents!)

Exercise 8.39. Let R be the ring of all C ∞ functions f : R → R.


Let m = {f ∈ R | f (0) = 0}.
a) Show that m = xR is a maximal ideal of R.
b) Show that forTall n ∈ Z+ , mn = {f ∈ R | f (0) = f 0 (0) = . . . = f (n−1) (0)}.

c) Deduce that n=1 mn is the ideal of all smooth T∞functions with identically zero
Taylor series expansion at x = 0. Conclude that n=1 mn 6= 0.
−1
d) Let f (x) = e x2 for x 6= 0 and 0 for x = 0. Show that f ∈/ f m.
e) Deduce that R is not Noetherian.
S∞ 1
Exercise 8.40. Let R = n=1 C[[t n ]] be the Puiseux series ring. Show that
R is a domain with a unique maximal ideal m and that for all n ∈ Z+ , mn = m.
Deduce from the Krull Intersection Theorem that R is not Noetherian.
The preceding exercise will become much more routine when we study valuation
rings in §17. In that language, one can show that if RTis a valuation ring with

divisible value group, then (R, m) is a local domain and n=1 mn = m.
184 8. NOETHERIAN RINGS

Exercise 8.41 (Kearnes-Oman [KeOm10]). Let I be an ideal in a Noetherian


ring R, and suppose that either R is a domain and I is proper or I is contained in
the Jacobson radical of
QR. Show: #R ≤ (#R/I)ℵ0 .

(Suggestion: Show # n=1 R/I n ≤ (#R/I)ℵ0 and apply Krull Intersection.)

13. Krull’s Principal Ideal Theorem


Theorem 8.44. (The Principal Ideal Theorem, a.k.a. Krull’s Hauptidealsatz)
Let x be a nonunit in a Noetherian ring R, and let p be minimal among prime ideals
containing x. Then p has height at most one.
Remark: A prime p which is minimal among primes containing x will be called a
minimal prime over x. Note that an equivalent condition is that p is a minimal
prime in the quotient ring R/(x). Note also that if x is nilpotent, every prime of p
contains x so the height of any minimal prime is 0.

Our strategy of proof follows Kaplansky, who follows D. Rees. We need a pre-
liminary result:
Lemma 8.45. Let u and y be nonzero elements in a domain R. Then:
a) The R-modules hu, yi/(u) and hu2 , uyi/(u2 ) are isomorphic.
b) If we assume further that for all t ∈ R, tu2 ∈ (y) implies tu ∈ (y), then the
R-modules (u)/(u2 ) and hu2 , yi/hu2 , uyi are isomorphic.
Proof. a) The isomorphism is simply induced by multiplication by u.
b) The module (u)/(u2 ) is cyclic with annihilator (u), and conversely any such
module is isomorphic to R/(u). Moreover M := hu2 , yi/hu2 , uyi is also cyclic,
being generated simply by y. Certainly u annihilates M , so it suffices to show that
the annihilator is exactly (u). More concretely, given ky = au2 + buy, we must
deduce that k ∈ (u). But we certainly have au2 ∈ (y), so by hypothesis au ∈ (y),
say au = cy. Then ky = cuy + buy. Since 0 6= y in our domain R, we may cancel y
to get k = (c + b)u ∈ (u). 
Proof of Krull’s Hauptidealsatz : Under the given hypotheses, assume for a contra-
diction that we have
p2 ( p1 ( p.
Note first that we can safely pass to the quotient R/p2 and thus assume that R is
a domain. Dually, it does not hurt any to localize at p. Therefore we may assume
that we have a Noetherian local domain R with maximal ideal m, an element x ∈ m,
and a nonzero prime ideal, say p, with x ∈ p ( m, and our task is now to show
that this setup is impossible. Now for the clever part: let 0 6= y be any element of
p, and for k ∈ Z+ , let Ik denote the ideal of all elements t with txk ∈ (y). Then
{Ik }∞
k=1 is an ascending chain of ideals in the Noetherian ring R so must stabilize,
say at k = n. In particular, tx2n ∈ (y) implies txn ∈ (y). Putting u = xn , we have
tu2 ∈ (y) implies (tu) ∈ (y).
Since m is a minimal prime over (x), the quotient ring T = R/(u2 ) has exactly
one prime ideal, m, and is therefore, by the Akizuki-Hopkins Theorem, an Artinian
ring, so that any finitely generated T -module has finite length. In particular, M :=
hu, yi/(u2 ), which can naturally be viewed as a T -module, has finite length, and
hence so does its T -submodule M 0 := hu2 , yi/(u2 ). Put N = hu2 , yihu2 , uyi. Then
`(M 0 ) = `(N ) + `(hu2 , uy/(u2 )) = `((u)/(u2 )) + `(hu, yi/(u)) = `(M );
13. KRULL’S PRINCIPAL IDEAL THEOREM 185

in the second equality we have used Lemma 8.45. The only way that M could have
the same length as its submodule M 0 is if hu, yi = hu2 , yi, i.e., if there exist c, d ∈ R
such that u = cu2 + dy, or u(1 − cu) = −dy. Since u lies in the maximal ideal of
the local ring R, 1 − cu ∈ R× , and thus u ∈ (y) ⊂ p. But m is minimal over x and
hence, being prime, also minimal over u = xn , contradiction! 
Corollary 8.46. With hypotheses as in Theorem 8.44, suppose that x is not
a zero-divisor. Then any prime p which is minimal over x has height one.
Exercise 8.42. Use the Akizuki-Hopkins theorem and Proposition 8.36 to give
a proof of Corollary 8.46.
Again we need a small preliminary result.
Lemma 8.47. (Prime Avoidance) Let R be a ring, and I1 , . . . , In , J be
Snideals
of R. Suppose that all but at most two of the Ii ’s are prime and that J ⊂ i=1 Ii .
Then J ⊂ Ii for some i.
Proof. We go by induction on n, the case n = 1 being trivial.
n = 2: Seeking a contradiction, suppose there is x1 ∈ J \ I2 and x2 ∈ J \ I1 . Since
J ⊂ I1 ∪ I2 we must have x1 ∈ I1 and x2 ∈ I2 . Then x1 + x2 ∈ J ⊂ I1 ∪ I2 .
If x1 + x2 ∈ I1 , then since x1 + x2 , x1 ∈ I1 , so is x2 , contradiction; whereas if
x1 + x2 ∈ I2 , then since x1 + x2 , x2 ∈ I1 , so is x1 .3
n ≥ 3: We may suppose S that In is prime and also that for all proper subsets
S ⊂ {1, . . . , n}, J 6⊂ i∈SSIi ; otherwise we would be done by induction. So for
1 ≤ i ≤ n, there is xi ∈ J \ j6=i Ij , and then xi ∈ Ii . Consider x = x1 · · · xn−1 +xn .
Then x ∈ J, so x ∈ Ii for some i.
Case 1: x ∈ In . Then since xn ∈ In , x1 · xn−1 ∈ In , and since In is prime xi ∈ In
for some 1 ≤ i ≤ n − 1, contradiction.
Case 2: x ∈ Ij for some 1 ≤ j ≤ n − 1. Then x1 · · · xn−1 ∈ Ij , so xn ∈ Ij ,
contradiction. 
Exercise 8.43. ([CDVM13, Prop. 2.2]) Let R be a UFD and not a field.
Suppose R× is finite. Show: R has infinitely many principal prime ideals.
(hint: Suppose R has finitely many principal nonzero principal prime ideals, say
(π1 ), . . . , (πn ). Let m ∈ MaxSpec
Sn R. By choosing x ∈ m• and applying unique
factorization, show m ⊂ i=1 (πi ). Apply Prime Avoidance and then Theorem
4.22.)
We can now give a striking structural result about primes in a Noetherian ring.
First a piece of notation: for any elements x, y in a partially ordered set S we
define the “interval” (x, y) to be the set of all z ∈ S such that x < z < y. For prime
ideals p and q, we denote by (p, q) the set of all prime ideals P with p ⊂ P ⊂ q.
Corollary 8.48. Let p ⊂ q be prime ideals in a Noetherian ring R. Then the
interval (p, q) is either empty or infinite.
Proof. As usual, by correspondence we may pass to R/p and therefore assume
that p = 0. Suppose that for some n ≥ 1 we haveS(0, q) = {p1 , . . . , pn }. BySPrime
n n
Avoidance (Lemma 8.47) we cannot then have q ⊂ i=1 pi , so choose x ∈ q\ i=1 pi .
Then q is a prime of R, of height at least 2, which is minimal over (x), contradicting
Theorem 8.44. 
3In fact this works for any subgroups I , I , J of a group G with J ⊂ I ∪ I .
1 2 1 2
186 8. NOETHERIAN RINGS

In particular, if R is Noetherian and Spec R is finite, then dim R ≤ 1.


Theorem 8.49. (Generalized Principal Ideal Theorem) Let R be a Noetherian
ring, and let I = ha1 , . . . , an i be a proper ideal of R. Let p be a minimal element
of the set of all prime ideals containing I. Then p has height at most n.
Proof. As usual, we may localize at p and suppose that R is local with p as
its maximal ideal. Suppose to the contrary that there exists a chain p = p0 ) p1 )
. . . ) pn+1 . Because R is Noetherian, we may arrange for (p1 , p) = ∅. Because p
is minimal over I, I cannot be contained in p1 ; without loss of generality we may
suppose that a1 is not in p1 . Put J := hp1 , a1 i; then J strictly contains p1 so p is
the unique prime of R containing J. So the ring R/J is an Artin local ring, and
then by Proposition 8.36 for sufficiently large k we have pk ⊂ J. Then by taking t
to be sufficiently large we can write, for 2 ≤ i ≤ n,
ati = ci a1 + bi , ci ∈ R, bi ∈ p1 .
Put K = hb2 , . . . , bn i ⊂ p1 . Since the height of p1 exceeds n − 1, by induction on n
we may assume that p1 properly contains a prime ideal Q which contains J. The
ideal Q0 := ha1 , Qi contains some power of each ai and therefore p is the unique
prime ideal containing Q0 . So in the quotient R/Q, the prime p/Q is minimal over
the principal ideal Q0 /Q. By Krull’s Hauptidealsatz (Theorem 8.44) p/Q has height
1. On the other hand, we have p/Q ) p1 /Q ) 0, a contradiction. 

14. The Dimension Theorem, following [BMRH]


The following is a very basic theorem about Noetherian rings:
Theorem 8.50. (Dimension Theorem) Let R be a Noetherian ring.
a) We have dim R[t] = dim R + 1.
b) We have dim R[[t]] = dim R + 1.
Remark: For a non-Noetherian ring R, one has the inequalities
dim R + 1 ≤ dim R[t] ≤ 2 dim R[t] + 1,
and indeed there are rings for which dim R[t] = 2 dim R + 1.

Of course, an immediate induction argument gives:


Corollary 8.51. Let k be a field. Then
dim k[t1 , . . . , tn ] = dim k[[t1 , . . . , tn ]] = n.
Traditional proofs of the Dimension Theorem require significant development of the
dimension theory of commutative rings, a topic which is not covered here. Hap-
pily, a striking alternate approach to the Dimension Theorem was given by Brewer,
Heinzer, Montgomery and Rutter in [BMRH]. We follow their treatment here.

COMPLETE ME! ♣

15. The Artin-Tate Lemma


Theorem 8.52. (Artin-Tate [AT51]) Let R ⊂ T ⊂ S be a tower of rings with:
(i) R Noetherian,
(ii) S finitely generated as an R-algebra, and
15. THE ARTIN-TATE LEMMA 187

(iii) S finitely generated as a T -module.


Then T is finitely generated as an R-algebra.
Proof. Let x1 , . . . , xn be a set of generators for S as an R-algebra, and let
ω1 , . . . , ωm be a set of generators for S as a T -module. For all 1 ≤ i ≤ n, we may
write
X
(24) xi = aij ωj , aij ∈ T.
j

Similarly, for all 1 ≤ i, j ≤ m, we may write


X
(25) ωi ωj = bijk ωk , bijk ∈ T.
i,j,k

Let T0 be the R-subalgebra of T generated by the aij and bijk . Since T0 is a finitely
generated algebra over the Noetherian ring R, the ring T0 is Noetherian by the
Hilbert Basis Theorem. Each element of S may be expressed as a polynomial in
the xi ’s with R-coefficients. Making substitutions using (24) and then (25), we see
S is generated as a T0 -module by ω1 , . . . , ωm , and in particular that S is a finitely
generated T0 -module. Since T0 is Noetherian, the submodule T is also finitely
generated as a T0 -module. This immediately implies that T is finitely generated as
a T0 -algebra and then in turn that T is finitely generated as an R-algebra. 
CHAPTER 9

Boolean rings

1. First Properties
Let R be a not-necessarily-commutative-ring with the property that x2 = x for all
x ∈ R. Then (1 + 1) = (1 + 1)2 = 1 + 1 + 1 + 1, so 1 + 1 = 0. It follows that −x = x
for all x ∈ R. Moreover for any x, y ∈ R, (x + y) = (x + y)2 = x2 + xy + yx + y 2 =
x + y + xy + yx, so xy + yx = 0 or xy = yx. Therefore R is commutative.

We define a Boolean ring to be a ring R such that x2 = x for all x ∈ R. That is,
a ring is Boolean iff all of its elements are idempotents.
Exercise 9.1. Show: the group of units of a Boolean ring is trivial.
Exercise 9.2. a) Show: a quotient of a Boolean ring is Boolean.
b) Show: a subring of a Boolean ring is Boolean.
Exercise 9.3. Show: a Boolean ring is absolutely flat.

2. Boolean Algebras
A Boolean ring is an object of commutative algebra. It turns out that there is a com-
pletely equivalent class of structures of an order-theoretic nature, called Boolean
algebras. In some ways the concept of a Boolean algebra is more intuitive and
transparent – e.g., starting directly from the definition, it is perhaps easier to give
examples of Boolean algebras. Moreover it is not at all difficult to see how to pass
from a Boolean ring to a Boolean algebra and conversely.

A Boolean algebra is a certain very nice partially ordered set (B, ≤). Recall
that for any partially orderet set B and any subset S, we have the notion of the
supremum sup S and the infimum inf S. To define these it is convenient to extend
the inequality notation as follows: if S, T are subsets of B, we write
S<T
to mean that for all s ∈ S and t ∈ T , s < t, and similarly
S≤T
to mean that for all s ∈ S and t ∈ T , s ≤ t.
Then we say that z = sup S if S ≤ z and if w is any element of B with S ≤ w,
then z ≤ w. Similarly z = inf S if z ≤ S and if w is any element of B with w ≤ S
then w ≤ Z. For a given subset S, neither sup S nor inf S need exist, but if either
exists it it is plainly unique. In particular if sup ∅ exists, it is necessarily a bottom
element, called 0, and if inf ∅ exists, it is necessarily a top element called 0.
189
190 9. BOOLEAN RINGS

A partially ordered set (L, ≤) is called a lattice if for all x, y ∈ L, sup{x, y} and
inf{x, y} both exist. We give new notation for this: we write
x ∨ y := sup{x, y},
the join of x and y and
x ∧ y := inf{x, y},
the meet of x and y.
A lattice is said to be bounded if it contains a bottom element 0 and a top element
1: equivalently, sup S and inf S exist for every finite subset S.
Exercise 9.4. Let L be a lattice containing 0 and 1, and let x ∈ L. Then:
a) x ∨ 1 = 1,
b) x ∧ 1 = x,
c) x ∨ 0 = x,
d) x ∧ 0 = 0.
A lattice L is complemented if it has a bottom element 0, a top element 1, and
for each x ∈ L there exists y ∈ L such that x ∨ y = 1, x ∧ y = 0.

A lattice is distributive if ∀x, y, z ∈ L,


(x ∨ y) ∧ z = (x ∧ z) ∨ (y ∧ z),
(x ∧ y) ∨ z = (x ∨ z) ∧ (y ∨ z).
Proposition 9.1. Let L be a distributive complemented lattice. Then for all
x ∈ L, the complement of x is unique.
Proof. Suppose that y1 and y2 are both complements to x, so
x ∨ y1 = x ∨ y2 = 1, x ∧ y1 = x ∧ y2 = 0.
Then
y2 = 1 ∧ y2 = (x ∨ y1 ) ∧ y2 = (x ∧ y2 ) ∨ (y1 ∧ y2 ) = 0 ∨ (y1 ∧ y2 ) = y1 ∧ y2 ,
so y2 ≤ y1 . Reasoning similarly, we get y1 ≤ y2 , so y1 = y2 . 
By virtue of Proposition 9.1 we denote the complement of an element x in a dis-
tributive complemented lattice as x∗ .
Exercise 9.5. Show: for every element x of a distributive complemented lattice
we have (x∗ )∗ = x.
A Boolean algebra is a complemented distributive lattice with 0 6= 1.
Exercise 9.6. Show: DeMorgan’s Laws hold in any Boolean algebra B: for
all x, y ∈ B, we have
a) (x ∧ y)∗ = x∗ ∨ y ∗ and
b) (x ∨ y)∗ = x∗ ∧ y ∗ .
The shining example of a Boolean algebra is the powerset algebra 2S for a nonempty
set S. In the special case in which |S| = 1, we denote the corresponding Boolean
algebra (the unique totally ordered set on two elements) simply as 2.

Not every Boolean algebra is isomorphic to a power set Boolean algebra.


2. BOOLEAN ALGEBRAS 191

Example: Let S be a set, and let Z(S) ⊂ 2S be the collection of all finite and
cofinite subsets of S. It is easily checked that (Z(S), ⊂) ⊂ (2S , ⊂) is a sub-Boolean
algebra. However, |Z(S)| = |S|, so if |S| = ℵ0 , then Z(S) is not isomorphic to any
power set Boolean algebra.

Boolean algebras form a full subcategory of the category of partially ordered sets.
In other words, we define a morphism f : B → B 0 of Boolean algebras simply to be
an isotone (or order-preserving) map: ∀x, y ∈ B, x ≤ y =⇒ f (x) ≤ f (y). One can
(and sometimes does, e.g. for model-theoretic purposes) also axiomatize Boolean
algebras as a structure (B, ∨, ∧, ∗, 0, 1), the point being that x ≤ y iff x ∨ y = y iff
x ∧ y = x, so the partial ordering can be recovered from the wedge or the join.
Proposition 9.2. The category of Boolean rings is equivalent to the category
of Boolean algebras.
In other words, we can define a functor F from Boolean rings to Boolean algebras
and a functor G from Boolean algebras to Boolean rings such that for every Boolean
ring R, R is naturally isomorphic to G(F (R)) and for every Boolean algebra B, B
is naturally isomorphic to F (G(B)).

Let us sketch the basic construction, leaving the details to the reader. Suppose
first that R is a Boolean ring. Then we associate a Boolean algebra F (R) with the
same underlying set as R, endowed with the following operations: ∀x, y ∈ R,
x ∧ y = xy,

x∗ = 1 − x.
The join operation is then forced on us by DeMorgan’s Laws:
x ∨ y = (x∗ ∧ y ∗ )∗ = x + y − xy.
Exercise 9.7. Check that (F (R), ∧, ∨, ∗) is indeed a Boolean algebra, and that
the bottom element 0 in F (R) (resp. the top element 1) is indeed the additive
identity 0 (resp. the multiplicative identity 1).
Conversely, suppose that we have a Boolean algebra (B, ∧, ∨, ∗). Then we define a
Boolean ring G(B) on the same underlying set B, by taking
x + y := (x ∧ y ∗ ) ∨ (y ∧ x∗ )

xy := x ∧ y.
The addition operation corresponds to the Boolean operation “exclusive or” or, in
more set-theoretic language, symmetric difference x∆y.
Exercise 9.8. Check that (G(B), +, ·) is indeed a Boolean ring with additive
identity the bottom element 0 of B and multiplicative identity the top element 1 of
B.
Exercise 9.9. a) Let R be a Boolean ring. Show: the identity map 1R on R
is an isomorphism of Boolean rings R → G(F (R)).
b) Let B be a Boolean algebra. Show that the identity map 1B on B is an isomor-
phism of Boolean algebras B → G(F (B)).
192 9. BOOLEAN RINGS

Exercise 9.10. Let X be a nonempty set, let BX be the Boolean algebra of


subsets of X, partially ordered by inclusion. Show: the corresponding Boolean ring
may be identified with the ring 2X of all functions from X to F2 under pointwise
addition and multiplication.
Exercise 9.11. Show: every finite Boolean algebra is isomorphic to a powerset
algebra. Deduce: every finite Boolean ring R is isomorphic to the ring of binary
functions on a finite set of cardinality log2 (#R). In fact, try to show this both
on the Boolean algebra side and on the Boolean ring side. (Hint for the Boolean
ring side: use the decomposition into a direct product afforded by an idempotent
element.)
Exercise 9.12. Show: an arbitrary direct product of Boolean algebras (or,
equivalently, Boolean rings) is a Boolean algebra (or...).
As we saw above, cardinality considerations already show that not every Boolean
algebra is the power set Boolean algebra, and hence not every Boolean ring is the
full ring of binary functions on some set X. However, in view of results like Cayley’s
theorem in basic group theory, it is a reasonable guess that every Boolean algebra is
an algebra of sets, i.e., is a sub-Boolean algebra of a power set algebra. We proceed
to prove this important result on the Boolean ring side.

3. Ideal Theory in Boolean Rings


Proposition 9.3. Let R be a Boolean ring.
a) For all x ∈ N and all n ≥ 2, xn = x.
b) A Boolean ring is reduced, i.e., has no nonzero nilpotent elements.
c) Every ideal in a Boolean ring is a radical ideal.
Proof. a) The case n = 2 is the definition of a Boolean ring, so we may
assume n ≥ 3. Assume the result holds for all x ∈ R and all 2 ≤ k < n. Then
xn = xn−1 x = x · x = x.
b) If x ∈ R is such that xn = 0 for some positive integer n, then either n = 1
or n ≥ 2 and xn = x; either way x = 0.
c) Let I be an ideal in the Boolean ring R. Then I = rad(I) iff R/I is reduced, but
R/I is again a Boolean ring and part b) applies. 
The ring Z/2Z is of course a Boolean ring. It is also a field, hence certainly a local
ring and a domain. We will shortly see that it is unique among Boolean rings in
possessing either of the latter two properties.
Proposition 9.4. a) The only Boolean domain is Z/2Z.
b) Every prime ideal in a Boolean ring is maximal.
Proof. a) Let R be a Boolean domain, and let x be an element of R. Then
x(x − 1) = 0, so in a domain R this implies x = 0 or x = 1, so that R ∼ = Z/2Z.
b) If p is a prime ideal in the Boolean ring R, then R/p is a Boolean domain,
hence – by part a) – is simply Z/2Z. But this ring is a field, so p is maximal. 
Proposition 9.5. Let R be a local Boolean ring. Then R ∼
= Z/2Z.
Proof. Let m be the unique maximal ideal of R. By Proposition 9.4, m is
moreover the unique prime ideal of R. It follows from Proposition 4.12d) that
m = nil R is the set of all nilpotent elements, so by Proposition 9.3b) m = 0. Thus
R is a field and thus, by Proposition 9.4 must be Z/2Z. 
3. IDEAL THEORY IN BOOLEAN RINGS 193

Exercise 9.13. Let R be a Boolean ring, let I be an ideal of R.


a) Show that x, y ∈ I =⇒ x ∨ y ∈ I.
b) Show in fact that hx, yi = hx ∨ yi.
c) Deduce that any finitely generated ideal of R is principal.
Proposition 9.6. a) For a Boolean ring R, the following are equivalent:
(i) R is finite.
(ii) R is Noetherian.
(iii) R has finitely many maximal ideals.
b) If these equivalent conditions hold, then R ∼
= (Z/2Z)n .
Proof. (i) =⇒ (ii) is clear.
(ii) =⇒ (iii): Since R is Noetherian and prime ideals are maximal, by the Akizuki-
Hopkins Theorem (Theorem 8.35) R is Artinian, and then Theorem 8.37 gives that
R is a finite product of local Boolean rings.
(iii) =⇒ (i): If MaxSpec QR is finite, then by Exercise 8.32 the ring R/J(R) is
n
a finite product of fields i=1 ki . In a Boolean ring the Jacobson radical is the
intersection of all maximal ideals, which is the intersection of all primeQideals,
which is the set of nilpotent elements, which is (0). So J(R) = 0 and R ∼
n
= i=1 ki .

Each ki is a quotient of R, hence Boolean, so ki = Z/2Z. So R is finite.
In the courseQof proving the equivalence of the conditions we also showed that
they imply R ∼
n
= i=1 Z/2Z. 
Lemma 9.7. For an ideal m in a Boolean ring R, the following are equivalent:
(i) m is maximal.
(ii) For all x ∈ R, either x ∈ R or 1 − x ∈ R (and not both!).
Proof. (i) =⇒ (ii): Of course no proper ideal in any ring can contain both
x and 1 − x for then it would contain 1. To see that at least one must lie in m, it
is certainly no loss to assume that x is neither 0 nor 1, hence R = xR × (1 − x)R.
Recall that in a product R1 × R2 of rings, every ideal I is itself a product I1 × I2 ,
where Ii is an ideal of Ri . Then R/I ∼ = R1 /I1 × R2 /I2 . So I is prime iff R/I is a
domain iff either (i) I1 is prime in R1 and I2 = R2 or (ii) I1 = R1 and I2 is prime in
R2 . In particular, every maximal ideal of R1 × R2 contains either R1 or R2 : done.
(i) =⇒ (ii):1 We prove the contrapositive: if m is not maximal, there exists a
maximal ideal M properly containing m. Let x ∈ M \ m. Since M is proper, it does
not also contain 1 − x, hence neither does the smaller ideal m. 
Exercise 9.14. (Kernel of a homomorphism) Let f : R → F2 be a homomor-
phism of Boolean rings.
a) Show that Ker f is ∨-closed: if x, y ∈ Ker f , then x ∨ y ∈ Ker f .
b) Show that Ker f is downward-closed: if x ∈ Ker f and y ≤ x, then y ∈ Ker f .
c) Explain why parts a) and b) are equivalent to showing that Ker f is an ideal of
the Boolean ring R.
d) Show that Ker f is in fact a maximal ideal of R.
e) Conversely, for every maximal ideal m of R, show that R/m = F2 and thus the
quotient map q : R → R/m is a homomorphism from R to F2 .

1This direction holds in any ring.


194 9. BOOLEAN RINGS

Exercise 9.15. (Shell of a homomorphism) Let f : R → F2 be a homomor-


phism of Boolean rings. Define the shell Sh f to be f −1 (1).
a) Show that Sh f is wedge-closed: if x, y ∈ Sh f , so is x ∧ y.
b) Show that Sh f is upward-closed: if x ∈ Sh f and x ≤ y, then y ∈ Sh f .
c) A nonempty, proper subset of a Boolean algebra which is wedge-closed and
upward-closed is called a filter, so by parts a) and b) Sh f is a filter on B. Show
that in fact it is an ultrafilter on B, i.e., that it is not properly contained in any
other filter. (Suggestion: use Lemma 9.7.)
d) Show that every ultrafilter on B is the shell of a unique homomorphism of Boolean
algebras f : B → F2 .

4. The Stone Representation Theorem


Let R be a Boolean ring. We would like to find an embedding of R into a Boolean
ring of the form 2X . The key point of course, is to conjure up a suitable set X.
Can we find any clues in our prior work on Boolean rings?
indent Well, finite Boolean rings we understand: the
L proof of Proposition 9.6 gives
us that every finite Boolean ring is of the form i∈X Z/2Z, where the elements
of X correspond to the maximal ideals of R. The isomorphism i∈X Z/2Z ∼ = 2X
L
amounts to taking each element x of R and recording which of the maximal ideals
it lies in: namely, x lies in the ith maximal ideal mi of all elements having a zero
in the ith coordinate iff its image in the quotient R/mi = Z/2Z is equal to 0.

This motivates the following construction. For any Boolean ring, let M (R) be
the set of all maximal ideals of R, and define a map E : R → M (R) by letting E(x)
be the set of maximal ideals of R which do not contain x. This turns out to be
very fruitful:
Theorem 9.8. (Stone Representation Theorem) Let R be a Boolean ring and
M (R) the set of maximal ideals. The map E : R → 2M (R) which sends an element
x of R to the collection of all maximal ideals of R which do not contain x is an
injective homomorphism of Boolean rings. Therefore R is isomorphic to the Boolean
ring associated to the algebra of sets E(R) ⊂ 2M (R) .
In particular this shows that every Boolean algebra is an algebra of sets.
Proof. Step 1: We check that the map E is a homomorphism of Boolean
algebras. Above we saw E(0) = ∅; also E(1) = M (R). Also, for x, y ∈ R, E(xy) is
the set of maximal ideals which do not contain xy; since maximal ideals are prime
this is the set of maximal ideals which contain neither x nor y, i.e., E(x) ∩ E(y) =
E(x) · E(y). Finally, E(x) + E(y) = E(x) ∆ E(y) is the set of maximal ideals which
contain exactly one of x and y, whereas E(x + y) is the set of maximal ideals not
containing x + y. For m ∈ M (R), consider the following cases:
(i) x, y ∈ m. Then m is not in E(x) ∆ E(y). On the other hand x + y ∈ m, so m is
not in E(x + y).
(ii) Neither x nor y is in m. Certainly then m is not in E(x) ∆ E(y). On the other
hand, remembering that R/m ∼ = Z/2Z, both x and y map to 1 in the quotient, so
x + y maps to 1 + 1 = 0, i.e., x + y ∈ m, so m is not in E(x + y).
(iii) Exactly one of x and y lies in m. Then m ∈ E(x) ∆E(y) and as above, x + y
maps to 1 in R/m, so x + y is not in m and m ∈ E(x + y).
Step 2: We show that the map E is injective. In other words, suppose we have two
5. BOOLEAN SPACES 195

elements x and y of R such that a maximal ideal m of R contains x iff it contains


y. Then
\ \
(x) = rad(x) = m= m = rad(y) = (y).
m∈M (R) | x∈ m m∈M (R) | y∈ m

So there are a, b ∈ R with y = ax, x = by, and then


x = by = by 2 = xy = ax2 = ax = y. 
Let us comment a bit on the proof. Although Step 1 is longer, it is rather routine.
They key is that E gives an embedding, which as we saw is equivalent to the
much gutsier statement that an element of a Boolean ring is entirely determined
by the family of maximal ideals containing it. From the standpoint of the more
“conventional” rings one encounters in number theory and algebraic geometry, this
is a very strange phenomenon. First of all, it can only be true in a ring R which
has trivial unit group, since if u is a nontrivial unit of course we will not be able to
distinguish 1 and u using ideals! Moreover it implies that every principal ideal is
radical, which does not hold in any domain which is not a field. Finally it implies
that every radical ideal is the intersection of the maximal ideals which contain it,
a property which we will meet later on: these are the Jacobson rings.

5. Boolean Spaces
We will now digress a bit to talk (not for the first or last time!) about topological
spaces. Following Bourbaki, for us compact means quasi-compact and Hausdorff.
Further a locally compact space is a Hausdorff space in which each point admits
a local base of compact neighborhoods. A subset of a topological space is clopen
if it is both closed and open.
A topological space X is totally disconnected if the only connected subsets of
X are the singleton sets {x}.2 A totally disconnected space is necessarily separated
(older terminology that I am not fond of: T1 ): i.e., singleton sets are closed. Indeed,
the closure of every connected set is connected, so the closure of a non-closed point
would give a connected set which is larger than a point. On the other hand a space
X is zero-dimensional if it admits a base of clopen sets.
Proposition 9.9. Let X be a locally compact space. Then X is totally discon-
nected iff it is zero-dimensional.
Proof. It is an exercise to show that every zero-dimensional Hausdorff space
is totally disconnected. For a proof that every locally compact totally disconnected
space is zero-dimensional, see e.g. [Cl-GT, Thm. 5.48]. 
A space X is called Boolean3 if it is compact and zero-dimensional; in particular
a Boolean space admits a base for the topology consisting of compact open sets.
Exercise 9.16. a) A finite space is Boolean iff it is discrete.
b) A Boolean space is discrete iff it is finite.
c) An arbitrary direct product of Boolean spaces is Boolean.
d) The usual Cantor space is homeomorphic to a countably infinite direct product
of copes of a discrete, two-point space and thus is a Boolean space.
2Following Qiaochu Yuan, we take the convention that the empty space is not connected: it
has zero connected components, not one!
3There are many synonyms: e.g. Stone space, profinite space.
196 9. BOOLEAN RINGS

Exercise 9.17. Show: a topological space is Boolean iff it is homeomorphic to


an inverse limit of finite, discrete spaces.
To every topological space X we may associate a Boolean algebra: namely, the
subalgebra of 2X consisting of compact open subsets. Thus in particular we may
associate a Boolean ring, say C(X), the characteristic ring of X. (We allow
ourselves to pass between C(X) and the associated Boolean algebra on the same
set and call the latter the characteristic algebra.)
Exercise 9.18. Show: the assignment X 7→ C extends to a contravariant func-
tor from the category of topological spaces to the category of Boolean rings.
(In other words, show that a continuous map f : X → Y of topological spaces
induces a “pullback” ring homomorphism C(f ) : C(Y ) → C(X).)
If X is itself a Boolean space, then the characteristic algebra C(X) is indeed char-
acteristic of X in the following sense.
Proposition 9.10. Let X be a Boolean space, and let A be a Boolean algebra
of subsets of X which is also a base for the topology of X. Then A = C(X).
Proof. By hypothesis the elements of A are open sets in X. Moreover, since
A is closed under complementation, the elements are also closed. Thus A ⊂ C(X).
Conversely, suppose Y ∈ C(X). Since Y is open and A is a base for the topology
on X, for each y ∈ Y there is Ay ∈ A with y ∈ Ay ⊂ Y . Thus {Ay }y∈Y is an
open cover for Y . But Y is also closed in a S
compact space hence itself compact, so
n
we may extract a finite subcover, say Y = i=1 Ayi . Since A is a subalgebra, it is
closed under finite unions, so Y ∈ A. Thus C(X) ⊂ A. 
To every Boolean ring R we may associate a Boolean space: namely there is a
natural topology on M (R), the set of maximal ideals of R, with respect to which
M (R) is a Boolean space. This topology can be described in many ways.

First Approach: by the Stone Representation Theorem we have an embedding


R ,→ 2M (R) and thus every element x ∈ R determines a function x : M (R) → F2 .
We may endow F2 with the discrete topology (what else?) and then give M (R)
the initial topology for the family of maps {x : M (R) → F2 }x∈R : the coarsest
topology which makes each of these maps continuous.

Here is a more concrete description of this initial topology: for each x ∈ R, put
Ux = {m ∈ M (R) | x ∈
/ m}
Vx = {m ∈ M (R) | x ∈ m}.
Then the topology in question is the one generated by {Ux , Vx }x∈X .

Second Approach: for any Boolean ring R, to give a maximal ideal m of R is


equivalent to giving a homomorphism of Boolean rings f : R → F2 . Namely, to a
maximal ideal m we associate the quotient map, and to f : R → F2 we associate the
kernel f −1 (0). In this way we get an embedding ι : M (R) ,→ 2R . Now we endow
each copy of F2 with the discrete topology and 2R with the product topology: this
makes it into a Boolean space.
Lemma 9.11. The image ι(M (R)) of ι is a closed subspace of 2R .
6. STONE DUALITY 197

Exercise 9.19. Prove Lemma 9.11.


Thus if we endow M (R) with the topology it inherits via the embedding ι, it is
itself a Boolean space.
Exercise 9.20. Show: the topology on M (R) defined via Lemma 9.11 coincides
with the initial topology on M (R) defined above.
It turns out to be important to consider a distinguished base for the topology on
M (R), which we now define. Since for a prime ideal m we have xy ∈ / m ⇐⇒ x ∈ /
m andy ∈ / m, we have for all x, y ∈ R that
Ux ∩ Uy = Uxy .
Moreover, by Lemma 9.7, M (R) \ Vx = Ux . It follows that the {Ux }x∈X form a
base of clopen sets for the topology on M (R).

To show the utility of this base, let us use it to show directly that M (R) is a
Boolean space.

Hausdorff: Let m1 and m2 be distinct maximal ideals of R. Choose x ∈ m2 \ m1 , so


by Lemma 9.7 1 − x ∈ m1 \ m2 . Thus m1 ∈ Ux m2 ∈ U1−x and
Ux ∩ U1−x = Ux(1−x) = U0 = ∅,
so we have separated m1 and m2 by open sets.

Quasi-compact: As is well-known, it is enough to check quasi-compactness of a


space using covers by elements of any fixed base. We certainly have a preferred
base here, namelyS{Ux }x∈X , so let’s use it: suppose that we have a collection
{xi }i∈I such that i∈I Uxi = M (R). Now again (and not for the last...) we exploit
the power of DeMorgan:
[ [ \
M (R) = Uxi = (M (R) \ Vxi ) = M (R) \ Vxi ,
i i i
T
so that i Vxi = ∅. This means that there is no maximal ideal containing every
xi . But that means that the ideal generated
Pby the xi ’s contains 1:Tthere exists a
finite subset J ⊂ I and aj ∈ R such that j aj xj = 1, and thus j∈J Vxj = ∅:
S
equivalently j∈J Uxj = M (R).

Thus we have shown that the correspondence R 7→ M (R) associates to every


Boolean ring a Boolean topological space, its Stone space.
Exercise 9.21. Show: the assignment R 7→ M (R) extends to a functor from
the category of Boolean rings to the category of Boolean spaces.

6. Stone Duality
Theorem 9.12. (Stone Duality) The functors C and M give a duality between
the category of Boolean spaces and the category of Boolean algebras. That is:
a) For every Boolean algebra B, the map B → C(M (B)) given by x ∈ B 7→ Ux is
an isomorphism of Boolean algebras.
b) For every Boolean space X, the map m : X → M (C(X)) given by x ∈ X 7→
mx := {U ∈ C(X) | x ∈ / U } is a homeomorphism of Boolean spaces.
198 9. BOOLEAN RINGS

Proof. a) The map e : x ∈ B 7→ Ux ∈ 2M (B) is the embedding e of the Stone


Representation Theorem. In particular it is an embedding of Boolean algebras. Its
image e(B) is a subalgebra of the characteristic algebra of the Boolean space M (B)
which is, by definition, a base for the topology of M (B). By Proposition 9.10 we
have e(B) = C(M (B)) so e is an isomorphism of Boolean algebras.
b) First we need to show that mx is a maximal ideal in the characteristic ring C(X).
It seems more natural to show this on the Boolean algebra side, i.e., to show that mx
is downward closed and union-closed. Indeed, U ∈ mx means x ∈ / U , so if V ⊂ X
then certainly x ∈ / V , i.e., V ∈ mx ; moreover, U, V ∈ mx ⇐⇒ x ∈ / U and x ∈ /
V ⇐⇒ x ∈ / U ∪ V ⇐⇒ U ∪ V ∈ mx . Thus mx is an ideal of C(X). Applying
Lemma 9.7, one easily sees that is maximal, so the map m is well-defined.
indent The injectivity of m follows immediately from the Hausdorff property of X.
Surjectivity: Let m ∈ M (C(X)). By Exercise X.X, we may identify m with
−1
a homomorphism of Boolean algebras fm : C(X) → F2 . Let F = fm (1) be the
shell of fm , an ultrafilter on the Boolean algebra of sets C(X). In particular F is
wedge-closed, i.e., it is a family of clopen subsets of the compact
T space X satisfying
the finite intersection property. Therefore there exists x ∈ U ∈F U . On the other
hand, the collection Fx of all clopen sets in X containing x is also a filter on
C(X) with F ⊂ Fx . But since F is an ultrafilter – i.e., a maximal filter – we have
F = Fx . Thus m and Fx are respectively the kernel and shell of the homomorphism
f : C(X) → F2 , so
m = C(X) \ Fx = {U ∈ C(X) | x ∈
/ U } = m(x).
Finally, since m is surjective, we have that for each A ∈ C(X),
{U ∈ M (C(X)) | A ∈ U } = {m(x) | x ∈ A},
so that m maps the base C(X) for the topology on X onto the base C(M (C(X))). 
Exercise 9.22. Let X be a topological space, and let C(X, 2) be the ring of all
continuous functions f : X → F2 (F2 being given the discrete topology).
a) Show: C(X, 2) is a Boolean ring.
b) Suppose X = M (R) is the maximal ideal space of the Boolean ring R. Show:
C(X, 2) is canonically isomorphic to R itself. Thus every Boolean ring is the ring
of continuous Boolean-valued functions on its Stone space of maximal ideals.
Exercise 9.23. Let S be a nonempty set and consider R = 2S = s∈S Z/2Z.
Q
a) Show that there is a natural bijective correspondence between elements of S and
principal maximal ideals of R.
b) Deduce that there is an embedding ι : S ,→ M (R) such that the induced topology
on S is discrete.
c) Show that ι(S) is dense in M (R). (Hint: R is atomic.)
d) Show that ι is a homeomorphism iff S is finite.
e)* Show that ι is the Stone-Cech compactification of the discrete space S.
CHAPTER 10

Associated Primes and Primary Decomposition

1. Associated Primes
Let M be an R-module. A prime ideal p of R is an associated prime of M if
there is m ∈ M with p = ann m = {x ∈ R | xm = 0}. The set of associated primes
of M is denoted (unfortunately) by Ass M .
Thus when R is a domain and M is torsionfree, (0) is the only associated prime
of M . In particular this holds for ideals of R. We hope this motivates the following
definition: for an ideal I of a ring R, the associated primes of the ideal I are the
associated primes of the module R/I.
Proposition 10.1.
For an R-module M and p ∈ Spec R, the following are equivalent:
(i) p ∈ Ass M .
(ii) There is an injection of R-modules R/p ,→ M .
Proof. (i) =⇒ (ii): Let p ∈ Ass M , and let m ∈ M with p = ann m. Define
ι : R → M by x 7→ xm. Then Ker ι = p, so ι gives an injection from R/p to M .
(ii) =⇒ (i): If ι : R/p ,→ M , let m = ι(1 + p). Then p = ann m. 
We immediately deduce:
Corollary 10.2. If N ⊂ M are R-modules, then Ass N ⊂ Ass M .
Proposition 10.3. For a prime ideal p of R, Ass R/p = {p}.
Proof. Proposition 10.1 gives p ∈ Ass R/p. Conversely, let x ∈ R be such that
ann(x+p) = q is a prime ideal. Since p is prime, y ∈ q ⇐⇒ yx ∈ p ⇐⇒ y ∈ p. 
For an R-module M , a zero divisor of M is an element x ∈ R such that xm = 0
for some m ∈ M • . We write ZD(M ) for the set of all zero divisors of M .
Proposition 10.4. For a nonzero R-module M , let F = {ann m | m ∈ M • }.
a) Every maximal element of F is a prime ideal.
b) If R is Noetherian, then Ass M 6= ∅.
Proof. a) Let I be an ideal of R of the form ann m for some x ∈ M • and not
properly contained in ann x0 for any x0 ∈ M • . Let a, b ∈ R be such that ab ∈ I
but b ∈/ I. Then bx ∈ M • . Since 0 = abx = a(bx), a ∈ ann(bx). But clearly
I = ann x ⊂ ann(bx), so by maximality of I we have I = ann(bx) and thus a ∈ I.
b) If M 6= 0, then F is a nonempty family of ideals in a Noetherian ring so has a
maximal element. Apply part a). 
Exercise 10.1. Let M be a nonzero R-module, and let S ⊂ R be a multiplica-
tive subset. Following [LR08, Prop. 3.5], we consider the family F of ideals I of R
with the following propery: for all m ∈ M , if Im = 0 then sm = 0 for some s ∈ S.
199
200 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

a) Show: F is increasing and strongly Ako.


b) Show: an ideal I of R is maximal with respect to being disjoint from S and
of the form ann(m) for some nonzero element m ∈ M iff it is a maximal
element of F 0 .
c) (Herstein) Show: an ideal that is maximal among annihilators of nonzero
points of M is prime.
d) Suppose that either R or M is Noetherian. Show: F is a monoidal filter.
Proposition
S 10.5. Let M be an R-module.
a) We have p∈Ass M p ⊂ ZD(M ).
S
b) If R is Noetherian, then p∈Ass M p = ZD(M ).
Proof. a) If p = ann m, then xm = 0 for all x ∈ p, so p ⊂ ZD(M ).
b) Let x ∈ ZD(M ), so that there is m ∈ M • with xm = 0. By Proposition 10.4
applied to N = hmi, there is p ∈ Ass N , i.e., there is y ∈ R such that ym 6= 0 and
p = ann ym. SSince xm = 0, xym = 0 and x ∈ p. By Proposition 10.2 p ∈ Ass M
and thus x ∈ p∈Ass M p. 

Proposition 10.6. Let N ⊂ M be R-modules. Then:


L⊂ Ass N
a) Ass M  ∪SAss M/N .
b) Ass i∈I M i = i∈I Ass Mi .

Proof. a) For p ∈ Ass M , let ι : R/p ⊂ M be an R-module monomorphism.


Put H = ι(R/p) and L = H ∩ N .
Case 1: Suppose L = 0. Then the natural map α : H → M/N is a monomorphism,
so α ◦ ι : R/p → M/N is a monomorphism and p ∈ Ass M/N .
Case 2: Let xL ∈ L• . Then x ∈ H • ∼= (R/p)• , so ann x = p.LSince x ∈SN , p ∈ Ass N .
b) Put M = i∈I Mi . Since each Mi is a submodule of i∈I MiS , i∈I Ass Mi ⊂
Ass M follows from Proposition 10.2. The containment Ass M ⊂ i∈I Ass Mi fol-
Lcase, let p ∈ Ass M . Then there is
lows from part a) when I is finite. In the general
an R-module monomorphism ι : R/p ,→ M = i∈ILMi . The image ι(R/p) lies in
the submodule generated by ι(1 + p), hence lies in i∈J Mi for some finite subset
J ⊂ I. This reduces us to the finite case. 

Theorem 10.7. Let R be a Noetherian ring, and let M be a nonzero, finitely


generated R-module.
a) There is a chain of submodules
0 = M0 ( M1 ( . . . ( Mn = M
such that for all 0 ≤ i ≤ n − 1 there is a prime ideal pi of R with Mi+1 /Mi ∼
= pi .
b) For any such chain, Ass M ⊂ {p1 , . . . , pn−1 }.
c) In particular, Ass M is finite.
Proof. a) By Proposition 10.5 M has an associated prime p1 = ann m1 . Put
M0 = {0} and M1 = hm1 i; note M1 /M0 = M1 ∼ = R/p1 . If M1 = M we’re
done; if not, M/M1 is finitely generated and nonzero so has an associated prime
p2 = ann(m2 + M1 ). Put M2 = hm1 , m2 i, so that M2 /M1 ∼ = R/p2 . We continue
in this way, getting an increasing chain of submodules Mi in M . Since M is
Noetherian, we must have Mn = M for some m.
b) By Proposition 10.3, for all 0 ≤ i ≤ n − 1 we have Ass Mi+1 /Mi = Ass R/pi =
{pi }. By Proposition 10.6 we have for all 0 ≤ i ≤ n−1, Ass Mi+1 ⊂ Ass Mi ∪{pi+1 },
1. ASSOCIATED PRIMES 201

and from this Ass M = Ass Mn ⊂ {p1 , . . . , pn } follows.


c) This follows immediately. 
Corollary 10.8. Let (R, m) be a Noetherian local ring. If m \ m2 consists
entirely of zero-divisors, then there is x ∈ R• with xm = 0.
Proof. If m = 0 we may take x = 1. Henceforth we assume m 6= 0, so by
Nakayama’s Lemma there is a ∈ m \ m2 . By SnTheorem 10.7 and Proposition 10.5,
Ass R = {p1 , . . . , pn } is finite and ZD(R) = i=1 pi . Thus by hypothesis
n
[
m \ m2 ⊂ pi .
i=1

For y ∈ m2 and p ∈ Z+ , a + y p ∈ m \ m2 , so by the Pigeonhole Principle there are


1 ≤ p < q ∈ Z+ such that a + y p , a + y q ∈ pi for some i. Then y p (1 − y q−p ) ∈ pi ;
since y q−p ∈ m and R is local, 1 − y q−p ∈ R× ; thus y p ∈ pi and, since pi is prime,
y ∈ pi . This shows
n
[
m⊂ pi .
i=1
By Prime Avoidance (Lemma 8.47), there is at least one i such that m ⊂ pi . By
definition pi = ann x for some x ∈ R• : we’re done. 
Proposition 10.9. Let S ⊂ R be multiplicatively closed.
a) If M is an S −1 R-module, then AssR M = AssS −1 R M .
b) If M is an R-module, then
AssR M ∩ Spec S −1 R ⊂ AssS −1 R S −1 M.
c) If R is Noetherian and M is an R-module, then
AssR M ∩ Spec S −1 R = AssS −1 R S −1 M.
Let M be an R-module. A weakly associated prime of M is a prime ideal p of
R such that there is x ∈ M with p = r(ann x). Thus the definition differs from the
usual one in that we are permitted to pass from ann x to its radical. We denote by
weakAss M the set of weakly associated primes of M .
Exercise 10.2. Show: for an R-module M and p ∈ Spec R, the following are
equivalent:
(i) p is weakly associated to M .
(ii) There is an ideal I of R with r(I) = p and an R-module injection R/I ,→ M .
Exercise 10.3. Show: parts a) and b) of Proposition 10.9b) hold if we replace
Ass by weakAss throughout.
Proposition 10.10. Let M be an R-module.
a) We have Ass M ⊂ weakAss M .
b) If R is Noetherian, then Ass M = weakAss M .
Proof. a) As the terminology suggests, this is immediate: if p = ann x, then
ann x is prime, hence radical, so p = r(ann x).
b) By Proposition 10.9 it is enough to show that p ∈ AssRp Mp : replacing R by Rp
we may assume R is Noetherian local with maximal ideal p. Since p ∈ weakAss M ,
there is x ∈ M with r(ann x) = p. Since R is Noetherian, by Proposition 4.15g) we
have that pn ⊂ ann x for some n ∈ Z+ . Again using the Noetherian hypothesis, the
202 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

set {ann y | y ∈ R is such that ann y ⊃ ann x} has a maximal element ann y, and
by Proposition 10.4, q = ann y is prime. Then we have pn ⊂ ann x ⊂ q, and since
q is prime and p is maximal, we have q = p and thus p ∈ Ass M . 
Exercise 10.4. Let k be a field and R = k[t1 , t2 , . . .] be the polynomial ring
in a countably infinite set of indeterminates over k. Let I = ht21 , t22 , . . .i, and let
p = r(I) = ht1 , t2 , . . .i. Show: p ∈ weakAss R/I \ Ass R/I.

2. The support of a module


For a module M over a ring R, we define its support
supp M = {p ∈ Spec R | Mp 6= 0}.
Proposition 10.11. For a finitely generated R-module M ,
supp M = {p ∈ Spec R | p ⊃ ann M }.
Proof. Write M = hω1 , . . . , ωn iR . For p ∈ Spec R, we have p ∈ supp M iff
Mp 6= 0 iff for some 1 ≤ i ≤ n we have ann(ωi ) ⊂ p. If these conditions hold then
n
\
ann M = ann(ωi ) ⊂ p.
i=1
Tn Qn
Conversely, if p contains ann(M ) = ann(ωi ) then p contains
i=1 i=1 ann(ωi )
hence it contains ann(ωi ) for some i, so p ∈ supp M . 
Theorem 10.12. Let M be an R-module.
a) We have weakAss M ⊂ supp M .
b) If R is Noetherian, the minimal elements of Ass M are precisely the minimal
elements of supp M .
c) The minimal associated primes of R are precisely the minimal primes of R.
Proof. a) Let p ∈ weakAss M . By Exercise X.X, there is an ideal I of R
with r(I) = p and an R-module embedding R/I ,→ M . Tensoring with the flat
R-module Rp gives an injection Rp /IRp ,→ Mp . Since r(I) = p, I ⊂ p and thus
IRp ⊂ pRp ( Rp and Mp ⊃ Rp /IRp 6= 0.
b) Recall that under the Noetherian assumption weakAss M = Ass M .
Step 1: Each p ∈ supp M contains an element of Ass M : indeed, let p ∈ supp M ,
so Mp 6= 0. Since Rp is Noetherian, Propositions 10.4b) and 10.9c) give
∅ 6= AssRp Mp = AssR M ∩ Spec Rp ,
and an element of the latter set is precisely an associated prime q of M with q ⊂ p.
Step 2: Let p ∈ Ass M be minimal, so by part a) p ∈ supp M . If there were
p0 ∈ supp M with p0 ( p then there is no element in Ass M which is contained in
p0 , contradicting Step 1.
Step 3: Let p ∈ supp M be minimal. By Step 1, p contains an element p0 of Ass M ,
but since Ass M ⊂ supp M and p is minimal we must have p = p0 .
c) Apply part b) to M = R. 
Theorem 10.13. If R is Noetherian, MinSpec R is finite.
Proof. Combine Theorem 10.7c) and Theorem 10.12c). 
Later we will give a topological proof of Theorem 10.13!
3. PRIMARY IDEALS 203

3. Primary Ideals
A proper ideal q of a ring R is primary if for all x, y ∈ R, xy ∈ q implies x ∈ q or
y n ∈ q for some n ∈ Z+ .
Exercise 10.5. a) Show: a prime ideal is primary. (Trivial but important!)
b) Show: an ideal q of R is primary iff every zero-divisor in R/q is nilpotent.
Neither the definition or primary ideal nor the characterization given in the above
exercise is particularly enlightening, so one natural question is: which ideals are
primary? (And, of course, another natural question is: what’s the significance of
a primary ideal?) Here are some simple results which give some information on
primary ideals, sufficient to determine all the primary ideals in some simple rings.
Proposition 10.14. Let q be an ideal in a ring R. If r(q) = m is a maximal
ideal, then q is primary. In particular, any power of a maximal ideal is primary.
Proof. Since r(q) is the intersection of all prime ideals containing q, if this
intersection is a maximal ideal m, then m is the unique prime ideal containing q and
R/q is a local ring with nil(R/q) = J(R/q) = m/q. In such a ring an element is a
zero-divisor iff it is a nonunit iff it is nilpotent, so q is primary. The “in particular”
follows since by Proposition 4.15f), r(mn ) = r(m) = m. 
Proposition 10.15. If q is a primary ideal, then its radical r(q) is a prime
ideal, the smallest prime ideal containing q.
Proof. Let xy ∈ r(q), so that (xy)m = xm y m ∈ p for some m ∈ Z+ . If xm
is in q then x ∈ r(q), so assume that xm is not in q. Then y m is a zero divisor in
R/q, so by definition of primary there exists n ∈ Z+ such that (y m )n ∈ q, and then
y ∈ r(q). The second statement holds for any ideal I whose radical is prime, since
r(I) is the intersection of all prime ideals containing I. 
A primary ideal is said to be p-primary if its radical is the prime ideal p.
Tn
Lemma 10.16. If q1 , . . . , qn are p-primary ideals, then q = i=1 qi is p-primary.
Proof. Let x, y be elements of the ring R such that xy ∈ q and xQ∈ R\q. Then
n
for all 1 ≤ i ≤ n, there exists ai ∈ Z+ such that y ai ∈ Ii , and then y i=1 ai ∈ q, so
q is primary. Moreover, by Proposition 4.15b),
\n n
\ \n
r(q) = r( qi ) = r(qi ) = q = q.
i=1 i=1 i=1

Exercise 10.6. Give an example of primary ideals q, q0 such that q ∩ q0 is not
primary.
Proposition 10.17. If q is a primary ideal, the quotient ring R/q is connected.
Proof. Indeed, a ring is disconnected if and only if it has an idempotent
element e different from 0 or 1. Such an element is certainly not nilpotent en = e
for all n – but is a zero-divisor, since e(1 − e) = e − e2 = 0. 
Exercise 10.7. Let k be a field, let R = k[x, y] and put I = (xy). Show: I is
not primary but “nevertheless” R/I is connected.
204 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

Example 10.18. We will find all primary ideals in the ring Z of integers.
Evidently (0) is prime and hence primary. If q is any nonzero primary ideal,
then its radical p = r(q) is a nonzero prime ideal, hence maximal. So, combining
Propositions 10.14 and 10.15 we find that a nonzero ideal in Z is primary iff its
radical is maximal. Moreover, for any prime power (pn ), r((pn )) = r((p)) = (p)
is maximal – we use here the elementary and (we hope) familiar fact that if p is a
prime number, (p) is a prime ideal (Euclid’s Lemma); such matters will be studied in
more generality in §X.X on factorization – so (pn ) is a primary ideal. Conversely,
if n is divisible by more than one prime power, then applying the Chinese Remainder
Theorem, we get that Z/n is disconnected.
Exercise 10.8. a) Let R be an domain for which each nonzero ideal is a (finite,
of course) product of maximal ideals. Use the above argument to show that an ideal
q of R is primary iff it is a prime power.
b) (For those who know something about PIDs) Deduce in particular that primary
= prime power in any principal ideal domain.
Consider the following property of an domain:

(DD) Every ideal can be expressed as a product of prime ideals.

This is a priori weaker than the hypothesis of Exericse X.Xa). Later we will devote
quite a lot of attention to the class of domains satisfying (DD), the Dedekind
domains. Among their many properties is that a Dedekind domain is (either a
field or) a domain in which each nonzero prime ideal is maximal. Thus in fact the
hypothesis of Exercise X.Xa) is equivalent to assuming that R is a Dedekind domain.

Remark(ably): Another characterization theorem says that any Noetherian domain


in which each primary ideal is a prime power is a Dedekind domain. In particu-
lar, any polynomial ring k[x1 , . . . , xn ] in 2 ≤ n < ∞ variables over a field admits
primary ideals which are not prime powers.

Exercise 10.9. Let R = Z[t]/(t2 + 3) (or, equivalently, Z[ −3]). Let q = (2).
a) Show that there is a unique ideal p2 with R/p2 = Z/2Z. Evidently p2 is maximal.
b) Show that r(q) = p2 , and deduce that I is primary.
c) Show that q is not a prime power, and indeed, cannot be expressed as a product
of prime ideals.
The ring R of Exercise 10.9 is a good one to keep in mind: it is simple enough to
be easy to calculate with, but it already displays some interesting general phenom-
ena. It is a Noetherian domain in which every nonzero prime ideal is maximal. It
is therefore “close” to being a Dedekind domain but it does not satisfy one other
property (“integral closure”) which will be studied later. It will turn out to be an
immediate consequence of the main result of this section that, notwithstanding the
fact that there are ideals which
√ do not factor into a product of primes, nevertheless
every proper ideal in R = Z[ −3] can be written as a product of primary ideals.
This, finally, is some clue that the notion of a primary ideal is a fruitful concept.
The following exercise gives an even simpler (and more explicit) example of a ring
R and a primary ideal q of R which is not a prime power.

Having seen examples of a primary ideals which are not prime powers, what about
4. PRIMARY DECOMPOSITION, LASKER AND NOETHER 205

the converse? Is it at any rate the case that any prime power is a primary ideal?
We know that this is indeed the case for powers of a maximal ideal. However, the
answer is again negative in general:

Example (Atiyah-MacDonald, p. 51): Let k be a field; put R = k[x, y, z]/(xy − z 2 ).


Denote by x, y, and z the images of x, y, z in R. Put p = hx, zi. Since R/p =
k[x, y, z]/(x, z, xy − z 2 ) = k[y] is a domain, p is a prime ideal. Now consider the
ideal p2 : we have xy = z 2 ∈ p2 , but x ∈ / p2 and y ∈
/ p = r(p2 ), so p2 is not primary.

4. Primary Decomposition, Lasker and Noether


Let R be a ring and I an ideal of R. A primary decomposition Tn of I is an ex-
pression of I as a finite intersection of primary ideals, say I = i=1 qi .

An ideal that admits at least one primary decomposition is said to be decom-


posable. This is not a piece of terminology that we will use often, but the reader
should be aware of its existence.

For any ring R, let us either agree that R itself admits the “empty” primary de-
composition or that R has no primary decomposition (i.e., it doesn’t matter either
way) and thereafter restrict our attention to proper ideals.

It may not be too surprising that not every ideal in every ring admits a primary
decomposition. Indeed, we will see later that if R is a ring for which (0) admits a
primary decomposition, then the ring R has finitely many minimal primes.

The first important result in this area was proved by Emanuel Lasker in 1905,
roughly in the middle of his 27 year reign as world chess champion. Here it is.
Theorem 10.19. (Lasker [La05]) Let R be a polynomial ring in finitely many
variables over a field. Every proper ideal I of R admits a primary decomposition.
Lasker’s proof of this theorem was a long and intricate calculation. As we will
shortly see, a broader perspective yields considerably more for considerably less
effort. In Lasker’s honor a ring R in which every proper ideal admits a primary
decomposition is called a Laskerian ring.
Exercise 10.10. If R is Laskerian and I is an ideal of R, then R/I is Laske-
rian.
Combining Lasker’s theorem with this exercise, we get that every finitely generated
algebra over a field admits a primary decomposition. This result is of fundamental
(indeed, foundational) importance in algebraic geometry.

However, in 1921 Lasker’s triumph was undeniably trumped by Emmy Noether.


To see how, we need one further concept. An ideal I is irreducible if whenever I
is written as an intersection of two ideals – i.e., I = J ∩ K – then I = J or I = K.
Exercise 10.11. Let I be a proper ideal in a principal ideal domain R. Show:
the following are equivalent:
(i) I is primary.
(ii) I is irreducible.
206 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

(iii) I is a prime power: there exists a in R and n ∈ Z+ such that (a) is prime and
I = (a)n = (an ).
Proposition 10.20. a) A prime ideal is irreducible.
b) An irreducible ideal in a Noetherian ring is primary.
Proof. a) Let p be a prime ideal, and write p = I ∩ J. Since then p ⊃ IJ, by
Proposition 4.9 we have p ⊃ I or p ⊃ J; WLOG say p ⊃ I. Then p = I ∩ J ⊂ I ⊂ p,
so that we must have I = p.
b) By passage to the quotient, we may assume that (0) is irreducible and show
that it is primary. So suppose xy = 0 and x 6= 0. Consider the chain of ideals
ann(y) ⊂ ann(y 2 ) ⊂ . . . ⊂ ann(y n ) ⊂ . . . .
Since R is Noetherian, this chain stabilizes: there exists n such that ann(y n ) =
ann(y n+k ) for all k. We claim that (x) ∩ (y n ) = 0. Indeed, if a ∈ (x) then
ay = 0, and if a ∈ (y n ) then a = by n for some b ∈ R, hence by n+1 = ay = 0, so
b ∈ ann(y n+1 ) = ann(y n ), hence a = by n = 0. Since the (0) ideal is irreducible, we
must then have y n = 0, and this shows that (0) is primary. 
Exercise 10.12. This exercise is taken from a post of E. Merkulova on
http://math.stackexchange.com/questions/28620. Let k be a field, R = k[x, y] and
I = hx2 , xy, y 2 i.
a) Show: I is primary. (Hint: use Proposition 10.14.)
b) Show: I = hx, y 2 i ∩ hx2 , yi.
c) Deduce: I is an ideal in a (very nice) Noetherian domain which is primary but
not irreducible.
Theorem 10.21. (Noether) Any proper ideal in a Noetherian ring admits a
primary decomposition.
Proof. Let I be a proper ideal in the Noetherian ring R. We claim I is a
finite intersection of irreducible ideals; in view of Proposition 10.20 this gives the
desired result. To see this: suppose that the set of proper ideals which cannot
be written as a finite intersection of irreducible ideals is nonempty, and choose a
maximal element I. Then I is reducible, so we may write I = J ∩ K where each of
J and K is strictly larger than I. But being strictly larger than I each of J and
K can be written as a finite intersection of irreducible ideals, and hence so can I.
Contradiction! 
In other words, a Noetherian ring is Laskerian. Therefore Lasker’s Theorem is
an immediate consequence of Noether’s Theorem together with the Hilbert Basis
Theorem, which we recall, was proved in 1888 and whose remarkably short and
simple – but nonconstructive – proof engendered first controversy and later deep
admiration. The same is true for Noether’s theorem: it is from this theorem, and
the ridiculous simplicity of its proof, that Noetherian rings get their name.
Exercise 10.13. Let R be a Noetherian
Tr ring, with minimal primes p1 , . . . , pr .
Show: there is N ∈ Z+ such that (0) = i=1 pN
i . (Hint: use Proposition 4.15g).)

5. Irredundant primary decompositions


If an ideal can be expressed as a product of prime ideals, that product is in fact
unique. We would like to have similar results for primary decomposition. Unfortu-
nately such a uniqueness result is clearly impossible. Indeed, if I = q1 ∩ . . . ∩ qn is
6. UNIQUENESS PROPERTIES OF PRIMARY DECOMPOSITION 207

a primary decomposition of I and p is any prime containing I, then q1 ∩ . . . ∩ qn ∩ p


is also a primary decomposition, and clearly a different one if p 6= qi for any i. A
proper ideal I may well be contained in infinitely many primes – e.g. by X.X this
occurs with I = (0) for any Noetherian domain of dimension at least 2 – so there
may well be infinitely many different primary decompositions.

But of course throwing in extra primes is both frivolous and wasteful. The fol-
lowing definition formalizes the idea of a primary decomposition which is “frugal”
in two reasonable ways.

A primary decomposition is said to be irredundant1 (or minimal, or reduced)


if both of the following properties hold:

(IPD1) For all i 6= j, r(qi ) 6= r(qj ). T


(IPD2) For all i, qi does not contain j6=i qj .

If wastefulness succeeds, so does frugality:


Lemma 10.22. An ideal that admits a primary decomposition admits an irre-
dundant primary decomposition.
Proof. By Lemma 10.16, we may replace any collection of primary ideals
qi with a common radical with their intersection and still have a primary ideal,
thus satisfying (IPD1). Then if (IPD2) is not satisfied, there is some qi which
contains the intersection of all the other qj ’s, hence it can be removed to obtain
a primary decomposition satisfying (IPD1) and with a smaller number of primary
ideals. Proceeding in this way we eventually arrive at an irredundant primary
decomposition. 
The question is now to what extent an irredundant primary decomposition is
unique. The situation here is significantly better: although the primary decom-
position is not in all cases unique, it turns out that there are some important
quantities which are defined in terms of a primary decomposition and which can
be shown to be independent of the choice of irredundant decomposition, i.e., are
invariants of the ideal. Such uniqueness results are pursued in the next section.

6. Uniqueness properties of primary decomposition


Recall that for ideals I and J of a ring R, (I : J) = {x ∈ R | xJ ⊂ I}, which is also
an ideal of R. We abbreviate (I : (x)) to (I : x) and ((x) : J) to (x : J).
Exercise 10.14. Show: for ideals I and J, I ⊂ (I : J).
Lemma 10.23. Let q be a p-primary ideal and x ∈ R.
a) If x ∈ q then (q : x) = R.
b) If x ∈
/ q then (q : x) is p-primary.
c) If x ∈
/ p then (q : x) = q.
Proof. a) If x ∈ q then 1(x) = x ⊂ q, so 1 ∈ (q : x).
/ q, so y n ∈ q for some n and thus
b) If y ∈ (q : x), then xy ∈ q; by assumption x ∈
1It is amusing to note that most dictionaries do not recognize “irredundant” as an English
word, but mathematicians have been using it in this and other contexts for many years.
208 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

y ∈ r(q) = p. So q ⊂ (q : x) ⊂ p; taking radicals we get r((q : x)) = p. Moreover, if


/ (q : x), then xyz = y(xz) ∈ q, so (xz)n = xn z n ∈ q for some
yz ∈ (q : x) with y ∈
n
n, and x ∈ / q =⇒ (z m )n ∈ q for some n ∈ Z+ , thus z mn ∈ q ⊂ (q : x).
c) We have in all cases that q ⊂ (q : x). If x ∈ / p = r(q) and y ∈ (q : x), then
xy ∈ q; since no power of x is q, we must have y ∈ q. 
Tn
Theorem 10.24. (First Uniqueness Theorem) Let I = i=1 qi be any irre-
dundant primary decomposition of the ideal I. Let pi = r(qi ). Then the pi ’s are
precisely the prime ideals of the form r((I : x)) as x ranges through elements of R.
In particular, they are independent of the choice of irredundant primary decompo-
sition.
T T
Proof. For x ∈ R we have (I : x) = ( i qi : x) = i (qi : x), so
\ \
r((I : x)) = r((qi : x)) = pj
i x∈q
/ j

by Lemma 10.23. If r(I : x) is prime, then r(I : x) =


pj for someTj. Conversely, for each i, by irredundancy of the decomposition there
exists xi ∈ j6=i qj \ qi and then the Lemma implies r(I : xi ) = pi . 
Corollary 10.25. Let R be Noetherian, and let I ( R be a proper ideal.
Let p1 , . . . , pr be the radicals of the primary ideals in an(y) irredundant primary
decomposition of I. Then
{p1 , . . . , pr } = Ass R/I.
Proof. By Theorem 10.24 the pi ’s are precisely the elements of weakAss R/I.
Since R is Noetherian, so is R/I and thus by Proposition 10.10b) weakAss R/I =
Ass R/I. 
Tn
Proposition 10.26. Let I = i=1 qi be a primary decomposition of an ideal
I, with pi = r(qi ). Then any prime ideal p containing I contains pi for some i.
T
Proof. If p ⊃ I = i qi , then
\ \
p = r(p) ⊃ r(qi ) = pi .
i i
Since p is prime, p ⊃ pi for some i. 

Exercise 10.15. Show: an infinite Boolean ring is not Laskerian.


Tn
Proposition 10.27. Let I ⊂ R be a decomposable ideal, I = i=1 qi an irre-
dundant primary decomposition, and pi = r(qi ). Then
n
[
pi = {x ∈ R : (I : x) 6= I}.
i=1
In particular, if the zero ideal is decomposable, then the set of zero divisors of R is
the union of the minimal associated primes of R.
Proof.
Tr By passage to the quotient ring R/I, we may assume that I = 0. Let
0 = i=1 qi be a primary decomposition, with pi = r(qi ). For x ∈ R, ((0) : x) 6= (0)
iff x is a zero-divisor, so it suffices to show the last statement of the proposition,
that the union of the minimal primes is the set of all zero-divisors. Let D be the
6. UNIQUENESS PROPERTIES OF PRIMARY DECOMPOSITION 209

set of all zero divisors, so from Exercise 3.X and the proof of Theorem 10.24 we
have [ [ \ [
D = r(D) = r((0 : x)) = pj ⊂ pj .
06=x 06=x x∈q
/ j j

Conversely, by Theorem 10.24 each pi is of the form r((0 : x)) for some x ∈ R. 
Theorem 10.28. (Second Uniqueness Theorem) Let I be an ideal of R, and
let
n
\ m
\
qi = I = rj
i=1 j=1
be two irredundant primary decompositions for an ideal I. By Theorem 10.24 we
know that m = n and that there is a reordering r1 , . . . , rn of the rj ’s such that for
1 ≤ i ≤ n, r(qi ) = pi = r(ri ). Moreover, if pi is minimal, then qi = rj .
In other words, the primary ideals corresponding to the minimal primes are inde-
pendent of the primary decomposition.

We will use the technique of localization to prove this result, so first we need
some preliminaries on the effect of localization on a primary decomposition.
Proposition 10.29. Let R be a ring, S ⊂ R a multiplicatively closed set, and
q be a p-primary ideal. Write ι : R → S −1 R for the localization map.
a) If S ∩ p 6= ∅, then ι∗ (q) = S −1 R.
b) If S ∩ p = ∅, then ι∗ (q) is ι∗ (p)-primary, and ι∗ (ι∗ (q)) = q.

Proof. a) If x ∈ S ∩ p, then for some n ∈ Z+ , xn ∈ S ∩ q, so ι∗ (q) contains a


unit of S −1 R and is therefore S −1 R. Part b) follows immediately from Proposition
7.3 and Proposition 7.5a). 

Tn Proposition 10.30. Let S ⊂ R be a multiplicatively closed set, and let I =


i=1 qi be an irredundant primary decomposition of an ideal I. Put pi = r(qi ) and
suppose that the numbering is such that S ∩ pi = ∅ for i ≤ m and S ∩ pi 6= ∅ for
i > m. Then:
m
\
ι∗ (I) = ι∗ (qi ),
i=1
m
\

ι ι∗ (I) = qi ,
i=1
and both of these are irredundant primary decompositions.
Exercise 10.16. Prove Proposition 10.30.
Proof of Theorem 10.28: let pi be a minimal associated prime, and put S = R \ pi .
Certainly S is a multiplicatively closed set, and moreover by minimality pi is the
unique associated prime which is disjoint from S. Applying Proposition 10.30 to
both primary decompositions gives
qi = ι∗ ι∗ (I) = ri .

210 10. ASSOCIATED PRIMES AND PRIMARY DECOMPOSITION

7. Applications in dimension zero


We now give the proof of the uniqueness portion of Theorem 8.37. Let m1 , . . . , mn
be the distinct maximal ideals of the Artinian
Qn ring R. As in the proof of Theorem
8.37a) there exists k ∈ Z+ such that i=1 mki = ∩ni=1 mki = 0. For each i, the
radical r(mki ) is the maximal k
Tnidealk mi , so by Proposition 10.14 each mi is an mi -
primary ideal. Thus 0 = i=1 mi is a primary decomposition of the zero ideal
which is moreover immediately seen to be irredundant. Since all the primes mi
are maximal, the desired uniqueness statement of Theorem 8.37b) follows from the
Second Uniqueness Theorem (Theorem 10.28) for primary decompositions.

8. Applications in dimension one


Let R be a one-dimensional Noetherian domain, and T I a nonzero ideal. Then by
n
Theorem 10.21, I has a primary decomposition: I = i=1 qi , where pi = r(qi ) ⊃
qi ⊃ I is a nonzero prime ideal. But therefore each pi is maximal, so that the pi ’s
are pairwise comaximal. By Proposition 4.18, so too are the qi ’s, so the Chinese
Remainder Theorem applies to give
\n Yn
I= qi = qi ,
i=1 i=1
and
n
R/I ∼
Y
= R/qi .
i=1
Thus in this case we can decompose any proper ideal as a finite product of primary
ideals and not just a finite intersection. Moreover, for I 6= 0, all the associated
primes are minimal over I, so the Uniqueness Theorems (Theorems 10.24 and 10.28)
simply assert that the ideals qi are unique. This observation will be very useful in
our later study of ideal theory in one dimensional Noetherian domains.
CHAPTER 11

Nullstellensätze

Let k be a field. By an affine algebra over k we simply mean a finitely generated


k-algebra. Of all the various and sundry classes of commutative rings we have met
and will meet later in these notes, affine algebras are probably the most important
and most heavily studied, because of their connection to algebraic geometry.

1. Zariski’s Lemma
In 1947 Oscar Zariski published a short note [Za47] proving the following result.
Theorem 11.1. (Zariski’s Lemma) Let k be a field, A a finitely generated
k-algebra, and m ∈ MaxSpec A. Then A/m is a finite degree field extension of k.
Exercise 11.1. Show: the following is an equivalent restatement of Zariski’s
Lemma: let K/k be a field extension such that K is finitely generated as a k-algebra.
Then K/k is an algebraic field extension.
Notwithstanding its innocuous appearance, Zariski’s Lemma is a useful result on
affine algebras over any field. Further, when k is algebraically closed, it carries all
of the content of Hilbert’s Nullstellensatz, the main theorem of this section.

So how do we prove Zariski’s Lemma?

Oh, let us count the ways! The literature contains many interesting proofs, employ-
ing an impressively wide range of ideas and prior technology. We will in fact give
several different proofs during the course of these notes. Of course some pride of
place goes to the first proof that we give, so after much thought (and after changing
our mind at least once!) we have decided on the following.
1.1. Proof of Zariski’s Lemma via the Artin-Tate Lemma.

As in Exercise 11.1, it suffices to prove the following: let K/k be a field exten-
sion that is finitely generated as a k-algebra. We claim K/k is algebraic.
Indeed, if not, let x1 , . . . , xn be a transcendence basis for K/k (n ≥ 1 since
K/k is transcendental), put k(x) = k(x1 , . . . , xn ) and consider the tower of rings
(26) k ⊂ k(x) ⊂ K.
To be sure, we recall the definition of a transcendence basis: the elements xi are alge-
braically independent over k and K/k(x) is algebraic. But since K is a finitely gen-
erated k-algebra, it is certainly a finitely generated k(x)-algebra and thus K/k(x)
is a finite degree field extension. Thus the Artin-Tate Lemma applies to (26): we
conclude that k(x)/k is a finitely generated k-algebra. But this is absurd. It implies
the much weaker statement that k(x) = k(x1 , . . . , xn−1 )(xn ) is finitely generated
211
212 11. NULLSTELLENSÄTZE

as a k(x1 , . . . , xn−1 )[xn ]-algebra, or weaker yet, that there exists some field F such
that F (t) is finitely generated as an F [t]-algebra: i.e., there exist finitely many ra-
tional functions {ri (t) = pqii(t)(t) N
}i=1 such that every rational function is a polynomial
in the ri ’s with k-coefficients. But F [t] is a PID with infinitely many nonassociate
nonzero prime elements q (e.g. adapt Euclid’s argument of the infinitude of the
primes), so we may choose a nonzero prime element q which does not divide qi (t)
for any i. It is then clear that 1q cannot be a polynomial in the ri (t)’s: for instance,
evaluation at a root of q in F leads to a contradiction. 

Remark: The phenomenon encountered in the endgame of the preceding proof


will be studied in great detail in §12. What we are actually showing is that for any
field F , the polynomial ring F [t] is not a Goldman domain, and indeed this is
closely related to the fact that Spec F [t] is infinite. More on this later.
1.2. McCabe’s Proof of Zariski’s Lemma.

We will give one further proof of Zariski’s Lemma now (and more later...), an
extremely elegant and simple one due to J. McCabe [McC76].

Let K/k be a field extension that is finitely generated as a K-algebra, say by


x1 , . . . , xn . Reorder the xi ’s so that x1 , . . . , xt are algebraically independent over
k and xt+1 , . . . , xn are algebraic over k(x1 , . . . , xn ). We may assume t ≥ 1, for
otherwise K/k is finitely generated algebraic field extension, hence of finite degree.
Let S = k[x1 , . . . , xt ], so S is a polynomial ring and is not a field. There is
y ∈ S • such that yxt+1 , . . . , yxn are all integral over S[ y1 ]. We have k ⊂ S[ y1 ] and
x1 , . . . , xt ∈ S[ y1 ], so K = k[x1 , . . . , xn ] is integral over S[ y1 ]. Since K is a field, by
Proposition 1.9 so is S[ y1 ].
Let m ∈ MaxSpec S. Since t ≥ 1, m 6= (0), so let f ∈ m• . Then f is invert-
ible in the field S[ y1 ] so there is g ∈ S and N ∈ Z+ such that f1 = ygN and thus
y N = f g. Since f ∈ m and maximal ideals are prime, y ∈ m. It follows that y
lies in every maximal ideal of S, hence 1 + y lies in no maximal ideal and is thus
a unit in S. But S × = k[x1 , . . . , xn ]× = k × , so 1 + y ∈ k × and y ∈ k • . Thus
k[x1 , . . . , xt ] = k[x1 , . . . , xt , y1 ] = S[ y1 ] is a field: contradiction!

2. Hilbert’s Nullstellensatz
Let k be a field, let Rn = k[t1 , . . . , tn ], and write An for k n . We introduce an anti-
tone Galois connection (V, I) between subsets of Rn and subsets of An . Namely:

For S ⊂ An , we put
I(S) = {f ∈ Rn | ∀x ∈ S, f (x) = 0}.
In other words, I(S) is the set of polynomials which vanish at every element of S.
Conversely, for J ⊂ Rn , we put
V (J) = {x ∈ An | ∀f ∈ J, f (x) = 0}.
This is nothing else than the Galois relation associated to the relation f (x) = 0 on
the Cartesian product Rn × An .
2. HILBERT’S NULLSTELLENSATZ 213

As usual, we would like to say something about the induced closure operators
on Rn and An . First, for any subset S of An , I(S) is not just a subset but an ideal
of Rn . In fact I(S) is a radical ideal: indeed, if f n ∈ I(S) then f n vanishes on
every point of S, so f vanishes at every point of S.

This little bit of structure pulled from thin air will quicken the heart of any Bour-
bakiste. But beyond the formalism, the key question is: exactly which sets are
closed? Without knowing this, we haven’t proved the Nullstellensatz any more
than the analogous formalities between sets and groups of automorphisms prove
the Galois correspondence for Galois field extensions.

Indeed, an ideal I is radical if f n ∈ I implies f ∈ I. But if f n vanishes iden-


tically on S, then so does f .

The closed subsets of An are closed under arbitrary intersections (including the
“empty intersection”: An = V ((0)) and under finite unions (including the “empty
union”: ∅ = V ({1}) = V (Rn ), and therefore form the closed sets for a unique
topology on An , the Zariski topology.
Exercise 11.2. a) Prove these facts.
b) Show that the Zariski topology on An/k coincides with the topology it inherits as
a subset of An/k .
c) Show that the Zariski topology is T1 : i.e., singleton subsets are closed.
d) Show that when n = 1, the Zariski topology is the coarsest T1 topology on k:
namely, the topology in which a proper subset is closed iff it is finite.
e) For any n ≥ 1, show that the Zariski topology on k n is discrete iff k is finite.
f ) For any infinite field and m, n ≥ 1, show that the Zariski topology on k m+n is
strictly finer than the product of the Zariski topologies on k m and k n .
Remark: It is often lamented that the Zariski topology (especially when k = C) is
so “coarse”. It is true that it is much coarser than the “analytic topology” on k n
when k is a topological field (i.e., the product topology from the topology on k).
But from an algebraic perspective the Zariski topology is if anything too fine: we
will see why later on when we extend the topology to all prime ideals on an affine
algebra. Moreover, the fact that the Zariski topology on Fnq is discrete creates many
geometric problems.
Exercise 11.3. Let k be a field and n ∈ Z+ as above. Explicitly compute the
ideal I(k n ), i.e., the set of all polynomials which vanish at every point of k n . Do
we necessarily have I(k n ) = {0}?
Lemma 11.2. For a = (a1 , . . . , an ) ∈ k n , put ma = hx1 − a1 , . . . , xn − an i.
Then:
a) We have Rn /ma = k. In particular ma is maximal.
b) ma = I({a}) is the ideal of all functions vanishing at a.
c) The assignment a 7→ ma is a bijection from k n to the set of all maximal ideals
m of Rn such that Rn /m = k.
Proof. Part a) is obvious (but important).
b) Certainly each xi − ai vanishes at a, so ma ⊂ I({a}). But by part a) ma is a
maximal ideal, whereas 1 ∈/ I({a}), so we must have ma = I({a}).
214 11. NULLSTELLENSÄTZE

c) The mapping a 7→ ma is an injection from k n to the set of maximal ideals with


residue field k. Conversely, let m be an ideal of Rn with Rn /m = k. For 1 ≤ i ≤ n
let ai be the image of xi in Rn /m = k. Then m ⊃ ma so we must have equality. 
We now pause for a very important definition. A ring R is a Jacobson ring if it
is “sufficiently many maximal ideals”: more precisely, such that every prime ideal
p of R is the intersection of the maximal ideals that contain it.
Exercise 11.4. a) Show: a ring R is a Jacobson ring iff for every ideal I, the
intersection of all maximal ideals containing I is rad(I).
b) Show: every homomorphic image of a Jacobson ring is Jacobson.
Proposition 11.3. (Rabinowitsch Trick [Ra30]) Let k be a field and n ∈ Z+ .
a) The ring R = k[x1 , . . . , xn ] is a Jacobson ring.
b) It follows that any affine k-algebra is a Jacobson ring.
Proof. a) It is sufficient to show that for each prime ideal p of R and a ∈ R\p,
there exists a maximal ideal m containing p and not containing a.
To show this, put Ra := R[ a1 ], and let pa = pRa be the pushed forward ideal.
Since p does not meet the multiplicative set generated by a, pa is still prime in
Ra . Let ma be any maximal ideal of Ra containing pa , and let m = ma ∩ R
be its contraction to R: a priori, this is a prime ideal. There is an induced k-
algebra embedding R/m ,→ Ra /ma . But Ra is still a finitely generated algebra
so by Zariski’s Lemma (Theorem 11.1) Ra /ma is finite dimensional as a k-vector
space, hence so is the subspace R/m. Thus the domain R/m must be a field: let
x ∈ (R/m)• , and write out a linear dependence relation of minimal degree among
the powers of x:
xn + cn−1 xn−1 + . . . + c1 x + c0 , ci ∈ k, c0 6= 0.
Thus
 −1
x xn−1 + cn−1 xn−2 + . . . + c1 = ,
c0
so x is invertible. Thus m is the desired maximal ideal.
b) This follows immediately from Exercise 11.4. 
Remark: It seems that the author “J.L. Rabinowitsch,” the author of [Ra30] is in
fact George Yuri Rainich, a distinguished Russian-American mathematician.

We prove one last fact before imposing the hypothesis that k is algebraically closed.
Proposition 11.4. Let k be any field and J an ideal of k[x] = k[x1 , . . . , xn ].
Then:
a) V (J) = V (rad J).
b) For any subset S ⊂ k n , I(S) is a radical ideal.
v) I(V (J)) is a radical ideal containing rad J.
Proof. The underlying mechanism here is the following truly basic observa-
tion: for f ∈ k[x], P ∈ k n and m ∈ Z+ , we have
f (P ) = 0 ⇐⇒ f m (P ) = 0.
a) Since J ⊂ rad J we have V (J) ⊃ V (rad J). Conversely, let P ∈ V (J) and
f ∈ rad J. Then there is m ∈ Z+ such that f m ∈ J, so f m (P ) = 0 and thus
f (P ) = 0. It follows that P ∈ V (rad J).
2. HILBERT’S NULLSTELLENSATZ 215

b) Similarly, for any f ∈ k[x] and m ∈ Z+ , if f m ∈ I(S), then for all P ∈ S,


f m (P ) = 0. But this implies f (P ) = 0 for all P ∈ S and thus f ∈ I(S).
c) This follows immediately from parts a) and b) and the tautological fact that for
any ideal J of k[x], I(V (J)) ⊃ J. 
Finally we specialize to the case in which the field k is algebraically closed. We have
done almost all the work necessary to establish the following fundamental result.
Theorem 11.5. (Hilbert’s Nullstellensatz) Let k be an algebraically closed
field, let k[x] = k[x1 , . . . , xn ]. Then:
a) I induces a bijective correspondence between the singleton sets of k n and the
maximal ideals: a ∈ k n 7→ ma = hx1 − a1 , . . . , xn − an i.
b) For any Zariski-closed subset S ⊂ k n , V (I(S)) = S.
c) For any ideal J of Rn , I(V (J)) = rad(J).
Thus there is an inclusion-reversing, bijective correspondence between Zariski-closed
subsets of k n and radical ideals of k[x].
Proof. a) Let m be a maximal ideal of k[x]. By Theorem 11.1, the residue
field k[x]/m is a finite degree extension of k. Since k is algebraically closed, this
forces k[x]/m = k, and now Lemma 11.2 applies.
b) There is no content here: it is part of the formalism of Galois connections.
c) By Proposition 11.4, it is no loss of generality to assume that J is a radical ideal.
Further, by Proposition 11.3, k[x] is a Jacobson ring, so any radical ideal J is the
intersection of the maximal ideals m containing it. This is true over any field k.
But combining with part a), we get that J is an intersection of maximal ideals of
the form ma for certain points a ∈ k n . Since ma = I({a}), J ⊂ ma iff every element
of J vanishes at a, in other words iff a ∈ V (J). Thus J is equal to the set of all
polynomials f ∈ Rn which vanish at every point of V (J): J = I(V (J))! 
Exercise 11.5. Let k be any field. Show that if either part a) or part c) of
Theorem 11.5 holds for the rings k[x1 , . . . , xn ], then k is algebraically closed. (Hint:
in fact both parts fail for each n ∈ Z+ , including n = 1.)
Exercise 11.6. Show: Zariski’s Lemma in the case that k is algebraically
closed is equivalent to the following statement: let J = hf1 , . . . , fm i be an ideal in
k[x1 , . . . , xn ]. Then either there exists a simultaneous zero a of f1 , . . . , fm or there
exist polynomials g1 , . . . , gm such that g1 f1 + . . . + gm fm = 1.
2.1. The Semirational Nullstellensatz.
Lemma 11.6. (Lang’s Lemma) Let k be a field, L an algebraically closed field,
ϕ : k → L a field embedding, and R a finitely generated k-algebra. Then there is a
homomorphism Φ : R → L extending ϕ.
Proof. Let m be a maximal ideal of k. By Zariski’s Lemma, R/m is a finite
degree field extension of k, so by basic field theory it embeds as a k-algebra into
any algebraically closed field containing k. 
Remark: In [Lg02, § IX.1], Lang gives a direct proof of Lemma 11.6. It is easy to
see that Lemma 11.6 implies Zariski’s Lemma, so this gives another way to proceed.
Corollary 11.7. Let k be a field, and let R be a domain that is finitely gener-
ated as a k-algebra. For any y1 , . . . , yn ∈ R• , there is a homomorphism ψ : R → k
such that ψ(yi ) 6= 0 for 1 ≤ i ≤ n.
216 11. NULLSTELLENSÄTZE

Proof. Apply Lang’s Lemma to the ring R[ y11 , . . . , y1n ]. 


n
Corollary 11.8. Let J be a proper ideal of k[t1 , . . . , tn ]. Then there is x ∈ k
such that for all f ∈ J, f (x) = 0.
Proof. Let m be a maximal ideal containing J. Zariski’s Lemma gives a k-
algebra embedding ψ : k[t1 , . . . , tn ]/m ,→ k. Let (x1 , . . . , xn ) = (ψ(t1 ), . . . , ψ(tn )).

n
For an ideal J of k[t1 , . . . , tn ], let V a (J) be the set of x = (x1 , . . . , xn ) ∈ k with
n
g(x) = 0 for all g ∈ J. For S ⊂ k , let I(S) be the set of g ∈ k[t1 , . . . , tn ] such that
g(x) = 0 for all x ∈ S. (Notice that we are extending to the algebraic closure on
the affine space side but not on the ring side, hence the term “semirational”.)
Theorem 11.9. (Semirational Nullstellensatz) For all ideals J of k[t1 , . . . , tn ],
we have I(V a (J)) = rad J.
Proof. It is easy to see that I(V a (J)) is a radical ideal containing J and
thus I(V a (J)) ⊃ rad J. Conversely, let f ∈ I(V a (J)). We must show that there
is N ∈ Z+ such that f m ∈ J. We may assume f 6= 0. We introduce a new
indeterminate tn+1 and let J 0 be the ideal hJ, 1 − tn+1 f i of k[t1 , . . . , tn , tn+1 ]. Let
n n+1
Z ⊂ k be the zero set of J, so f |Z ≡ 0. Let (x1 , . . . , xn , xn+1 ) ∈ k . If
0
(x1 , . . . , xn ) ∈
/ Z, then there is g ∈ J ⊂ J such that g(x1 , . . . , xn , xn+1 ) 6= 0.
If (x1 , . . . , xn ) ∈ Z, then 1 − xn+1 f (x1 , . . . , xn ) = 1 6= 0. By Corollary 11.8,
J 0 = k[t1 , . . . , tn , tn+1 ], so there are gi ∈ k[t1 , . . . , tn , tn+1 ] and hi ∈ J such that
1 = g0 (1 − tn+1 f ) + g1 h1 + . . . + gr hr .
Now substitute tn+1 = f −1 and multiply by an appropriate power f N of f to clear
denominators: we get f N ∈ J. 

Exercise 11.7. The argument used in the proof of Theorem 11.9 is also called
the Rabinowitsch Trick. Explain its relation to the proof of Proposition 11.3.
Exercise 11.8. Can you deduce Theorem 11.9 from Hilbert’s Nullstellensatz?
Exercise 11.9. Let R be a subring of an algebraically closed field k, and let
f1 , . . . , fr ∈ R[t1 , . . . , tn ]. Show that exactly one of the following holds:
(i) There is x ∈ k n such that f1 (x) = . . . = fr (x) = 0. Pr
(ii) There are g1 , . . . , gr ∈ R[t1 , ..., tn ] and a ∈ R• such that i=1 gi fi = a.
In the case of R = Z and k = C, Exercise 11.9 is often called the “Arithmetic
Nullstellensatz.” A more interesting version of it would ask in Case (ii) for explicit
upper bounds on the degrees of the polynomials gj .

3. The Real Nullstellensatz


Recall that a field k is formally real if it is not possible to express −1 as a sum
of (any finite number of) squares in k.
Exercise 11.10. Let k be a formally real field.
a) Show that k is not algebraically closed.
b) Show that any subfield of k is formally real.
3. THE REAL NULLSTELLENSATZ 217

A field k is real closed if it is formally real and admits no proper formally real al-
gebraic extension. So e.g. R is real closed and Q is formally real but not real closed.

As we saw, even the weak Nullstellensatz fails for polynomial rings over any non-
algebraically closed field k. However, when k is formally real one can find coun-
terexamples to the Nullstellensatz of a particular form, and when k is real closed
one can show that these counterexamples are in a certain precise sense the only
ones, leading to an identification of the closure operation J 7→ I(V (J)) in this case.

In any commutative ring R, an ideal I is real if for all n ∈ Z+ and x1 , . . . , xn ∈ R,


x21 + . . . + x2n ∈ I implies x1 , . . . , xn ∈ I.

A domain R is real if the zero ideal is real.


Exercise 11.11. a) Show: a domain is real iff its fraction field is formally
real.
b) Let R be a ring. Show that p ∈ Spec R is real iff the fraction field of R/p is
formally real.
Exercise 11.12. Show: any real ideal is a radical ideal.
So what? The following result gives the connection to the closure operator on ideals
in k[t1 , . . . , tn ].
Proposition 11.10. Let k be a formally real field and k[x] = k[x1 , . . . , xn ].
For any ideal J of k[x], its closure J = I(V (J)) is a real ideal.
Proof. Let f1 , . . . , fm ∈ k[x] be such that f12 + . . . + fm 2
∈ J. Then for any
2 2
P ∈ V (J), we have f1 (P ) + . . . + fm (P ) = 0. Since k is formally real, this implies
f1 (P ) = . . . = fm (P ) = 0, and thus f1 , . . . , fm ∈ I(V (J)) = J. 
Exercise 11.13. Find a real prime ideal p ∈ Q[t] which is not closed.
For an ideal I of a ring R, define the real radical
R ad(I) = {x ∈ R |∃n ∈ Z+ ∃b1 , . . . , bm ∈ R | x2n + b21 + . . . + b2m ∈ I}.
Proposition 11.11. [BCR, Prop. 4.1.7] Let I be an ideal in a ring R.
a) A real ideal J contains I iff J ⊃ R ad(I) i.e., R ad(I) is the unique minimal real
ideal containing I.
b) R ad(I) is equal to the intersection of all real prime ideals p ⊃ I.
c) It follows that every real ideal is equal to the intersection of all the real prime
ideals containing it.
Remark: If there are no real prime ideals containing I, then the intersection over
this empty set is taken to be R.
Proof. Step 1: we show that R ad(I) is an ideal. The only nonobvious part
of this is closure under addition. Suppose that
a2n + b21 + . . . + b2m , A2N + B12 + . . . + BM
2
∈ I.
We may write
(a + A)2(n+N ) + (a − A)2(n+N ) = a2m c + A2M C,
218 11. NULLSTELLENSÄTZE

with c, C sums of squares in R. Then


(a + A)2(n+N ) + (a − A)2(n+N ) + c(b21 + . . . + b2m ) + C(B12 + . . . + BM
2
) ∈ I,
so a + A ∈ R ad(I).
Step 2: R ad(I) is a real ideal. Indeed, if x21 + . . . + x2k ∈ R ad(I), then there exists
n ∈ Z+ and b1 , . . . , bm ∈ R such that
(x21 + . . . + x2k )2m + b21 + . . . + b2m ∈ I;
for each 1 ≤ i ≤ k, we may rewrite this as x4m i + B12 + . . . + BN
2
, so xi ∈ R ad(I).
Step 3: Since every real ideal is radical, it is clear that any real ideal containing I
also contains R ad(I).
Step 4: Let a ∈ R \ R ad(I). By Zorn’s Lemma, the set of real ideals containing I
but not a has a maximal element, say J. We claim that J is prime. If not, there
exist b, b0 ∈ R \ J such that bb0 ∈ J. Then a ∈ R ad(J + bR) and a ∈ R ad(J + b0 R),
hence there are j, j 0 ∈ J such that
0
a2m + c21 + . . . + c2q = j + bd, a2m + c02 02 0 0 0
1 + . . . + cq = j + b d .

It follows that
0
a2(m+m ) + a sum of squares = jj 0 + jb0 d0 + j 0 bd + bb0 dd0 ∈ J,
and thus a ∈ R ad(J) = J, contradiction. Thus R ad(I) is the intersection of all
real prime ideals containing I. 

Theorem 11.12. (Artin-Lang Homomorphism Theorem)


a) Let k ,→ L be a map of real-closed fields, and let R be a finitely generated k-
algebra. If there is a k-algebra homomorphism ϕ : R → L, then there is a k-algebra
homomorphism ψ : R → k.
b) Let k be a real-closed field, and let R be a domain which is a finitely generated
k-algebra. If R is real, there is a k-algebra homomorphism ϕ : R → k.
Proof. a) For a proof using model-theoretic methods, see e.g. [BCR, Thm.
4.1.2].
b) The fraction field K of R is formally real: let L be a real closure of K. Apply
part a) to the composite k-algebra homomorphism ϕ : R → K → L. 

Remark: For a direct algebraic proof of Theorem 11.12b), see e.g. [S, § 3.3].
Exercise 11.14. Let R = R[x, y]/(x2 + y 2 ).
a) Show that R is a domain.
b) Show that R is not real.
c) Show that there is a (unique!) R-algebra homomorphism ϕ : R → R.
(Thus the converse of Theorem 11.12b) does not hold.)
In [Lg53], Lang actually proved the following stronger result than Theorem 11.12b),
which we state in more geometric language: the domain R above corresponds to
an integral affine variety V over the real closed field k, and Lang showed that the
function field k(V ) is formally real iff V has a nonsingular k-rational point.
Theorem 11.13. (The Nullstellensatz for Real-Closed Fields) Let k be a real-
closed field, and J an ideal in k[t] = k[t1 , . . . , tn ]. Then J = R ad(J).
4. THE COMBINATORIAL NULLSTELLENSATZ 219

Proof. Step 1: Suppose J is a real prime ideal. Let R = k[t]/J, and let K be
the fraction field of R; by Exercise 11.X, K is formally real; let L be a real closure
1
of K. For f ∈ R \ J, let S be the localization R[ q(f ) ], so S ⊂ L. By Theorem
11.12a) there is a k-algebra homomorphism ψ : S → k; let x = (ψ(t1 ), . . . , ψ(tn )).
Then x ∈ V (J) and f (x) 6= 0, so f ∈/ I(V (J)). It follows that J = I(V (J)) = J.
Step 2: Suppose J is a real ideal, and let T XJ be the set of all real prime ideals
containing J. By Proposition 11.11c), J = p∈XJ p, and thus
\ [ \ \
J = I(V (J)) = I(V ( p)) = I( V (p)) = I(V (p)) = p = J.
p∈XJ p∈XJ p∈XJ p∈XJ

Step 3: Now let J be arbitrary. By Proposition 11.10, J is a real ideal containing


J, so by Proposition 11.11a), R ad(J) ⊂ J. On the other hand, part b) gives
J ⊂ R ad(J) = R ad(J),
so J = R ad(J). 

4. The Combinatorial Nullstellensatz


In this section we describe one of the most recent Nullstellensätze, a celebrated
result of Noga Alon that has served as a powerful technical tool and organizing
principle in combinatorics and additive number theory (!).
Lemma 11.14. (Alon-Tarsi) Let k be a field, n ∈ Z+ , and f (t) ∈ k[t] =
k[t1 , . . . , tn ]; for 1 ≤ i ≤ n,
Qnlet di be the ti -degree of f , let Si be a subset of k
with #Si > di , and let S = i=1 Si . If f (x) = 0 for all x ∈ S, then f = 0.
Proof. We go by induction on n. The case n = 1 is truly basic: a nonzero
univariate polynomial over a field has no more roots than its degree. Now suppose
n ≥ 2 and that the result holds for polynomials in n − 1 variables. The basic idea
is the identity k[t1 , . . . , tn−1 , tn ] = k[t1 , . . . , tn−1 ][tn ]: thus we write
tn
X
f= fi (t1 , . . . , tn−1 )tin
i=0

with fi ∈ k[t1 , . . . , tn−1 ]. If (x1 , . . . , xn−1 ) ∈ k n−1 , the polynomial f (x1 , . . . , xn−1 , tn ) ∈
k[tn ] vanishes for all #Sn > dn elements xn ∈ Sn and thus is identically zero, i.e.,
fi (x1 , . . . , xn−1 ) = 0 for all 0 ≤ i ≤ tn . By induction, each fi (t1 , . . . , tn−1 ) is the
zero polynomial and thus f is the zero polynomial. 
Theorem 11.15. (Combinatorial Nullstellensatz Qn [Al99]) Let k be a field, S1 , . . . , Sn
be nonempty finite subsets of k, and put S = i=1 Si . For 1 ≤ i ≤ n, put
Y
gi (ti ) = (ti − si ) ∈ k[ti ] ⊂ k[t] = k[t1 , . . . , tn ]
si ∈Si

and
g = hg1 (t1 ), . . . , gn (tn )i.
Then
S = V (g)
and
g = I(V (g)) = g.
220 11. NULLSTELLENSÄTZE

Proof. That S = V (g) is immediate, as is g ⊂ g. Conversely, suppose f ∈


g = I(S), i.e., f (s) = 0 for all s = (s1 , . . . , sn )P∈ S. We must show that there are
n
polynomials h1 (t), . . . , hn (t) such that f (t) = i=1 hi (t)gi (t).
For 1 ≤ i ≤ n, put di = #Si − 1; we may write
di
X
gi (ti ) = tdi i +1 − gij tji .
j=0

Observe that if si ∈ Si , then gi (xi ) = 0, i.e.,


di
X
(27) xdi i +1 = gij xji .
j=0

Let f be the polynomial obtained from f by writing f as a sum of monomials and


repeatedly substituting each instance of tei i with ei > di with a k-linear combination
Pn of ti using (27). Then f has degree at most di in ti and f − f is
of smaller powers
of the form i=1 hi gi . Further, for all s = (s1 , . . . , sn ) ∈ S, f (s) = f (s) = 0. Thus
Pn
Lemma 11.14 applies to give f = 0 and thus f = i=1 hi gi . 

Exercise 11.15. a) Show: in the notation of the proof of Theorem 11.13, the
polynomials h1 , . . . , hn satisfy deg hi ≤ deg f − deg gi for all 1 ≤ i ≤ n.
b) Show: the coefficients of h1 , . . . , hn lie in the subring of k generated by the
coefficients of f, g1 , . . . , gn .
Corollary 11.16. (Polynomial Method) Let k be a field, n ∈ Z+ , a1 , . . . , an ∈
N, and let f ∈ k[t] = k[t1 , . . . , tn ]. We suppose:
(i) deg f = a1 + . . . + an .
(ii) The coefficient of ta1 1 · · · tann in f is nonzero.
Then, for any subsets Q S1 , . . . , Sn of k with #Si > ai for 1 ≤ i ≤ n, there is
n
s = (s1 , . . . , sn ) ∈ S = i=1 Si such that f (s) 6= 0.
Proof. It is no loss of generality to assume that #Si = ai + 1 for all i, and
we do so. We will show that if (i) holds and f |S ≡ 0, then (ii) does not hold, i.e.,
the coefficient of ta1 1 · · · tann in f is 0. Q
Define, for all 1 ≤ i ≤ n, gi (ti ) = si ∈Si ti − si . By Theorem 11.13 and the
preceding exercise, there are h1 , . . . , hn ∈ k[t] such that
n
X
f= hi gi
i=1

and
deg hi ≤ (a1 + . . . + an ) − deg gi , ∀1 ≤ i ≤ n,
so
(28) deg hi gi ≤ deg f.
Thus if hi gi contains any monomial of degree deg f , such a monomial would be
of maximal degree in hi gi = hi si ∈Si (ti − si ) and thus be divisible by tai i +1 . It
Q
follows that for all i, the coefficient of ta1 1 · · · tann in hi gi is zero, hence the coefficient
of ta1 1 · · · tann in f is zero. 
5. THE FINITE FIELD NULLSTELLENSATZ 221

5. The Finite Field Nullstellensatz


For a prime power q, let Fq be a finite field of cardinality q. We will characterize
the closure operation J 7→ J = I(V (J)) for ideals in Fq [t] = Fq [t1 , . . . , tn ].

Let I0 = htq1 − t1 , . . . , tqn − tn i. Then the key observation is that for any ideal J of
Fq [t], J ⊃ I0 . Indeed, since xq = x for all x ∈ Fq , the polynomials tq1 −t1 , . . . , tqn −tn
each vanish at every point of Fnq , so J = I(V (J)) ⊃ I(Fnq ) ⊃ I0 . Since of course
J ⊃ J, it follows that for all ideals J of k[t] we have
(29) J ⊃ J + I0 .
Proposition 11.17. (Finite Field Weak Nullstellensatz) Let Fq be a finite
field, and let n ∈ Z+ . For 1 ≤ i ≤ n, let gi = tqi − ti ∈ Fq [t1 , . . . , tn ], and put
I0 = hg1 , . . . , gn i. Then I0 = I(Fnq ) is the ideal of all functions vanishing at every
point of Fnq .
Exercise 11.16. Deduce Proposition 11.17 from the Combinatorial Nullstel-
lensatz.
Proposition 11.17 asserts that the containment of (29) is an equality when J = (0).
In fact, this holds in all cases.
Lemma 11.18. Let J be an ideal of Fq [t1 , . . . , tn ]. If J contains the ideal I0 =
htq1 − t1 , . . . , tqn − tn i, then rad J = J.
Proof. Suppose that for x ∈ R, there is n ∈ Z+ with xn ∈ J. Then also
qn n−1
x = (xq )q ∈ J. By Corollary 11.17, for all x ∈ R, f q − f ∈ I0 ⊂ J: applying
n−1 n n n
this with f = xq , we find that xq − x ∈ I and thus xq − (xq − x) = x ∈ J. 
Theorem 11.19. (Finite Field Nullstellensatz) For any ideal J of R = Fq [t1 , . . . , tn ],
J = I(V (J)) = J + I0 = hJ, tq1 − t1 , . . . , tqn − tn i.
We will give two proofs: one using the Semirational Nullstellensatz, and one using
the Finite Field Weak Nullstellensatz.
Proof. By the Semirational Nullstellensatz (Theorem 11.9) and Lemma 11.18,
I(V a (J + I0 )) = rad(J + I0 ) = J + I0 .
Since V a (I0 ) = Fnq , we have
I(V (J)) = I(V a (J) ∩ Fnq ) = I(V a (J) ∩ V a (I0 )) = I(V a (J + I0 )) = J + I0 .

Proof. By (29) J ⊃ J + I0 , it suffices to show that for all J ⊇ I0 , J = J.
By Proposition 11.17, I0 = I(Fnq ). For P = (x1 , . . . , xn ) ∈ Fnq , let
mP = I({P }) = ht1 − x1 , . . . , tn − xn i.
T
Thus {mP }P ∈Fnq are finitely many pairwise comaximal ideals with I0 = P ∈Fn mP .
q
By the Chinese Remainder Theorem,
n
mP ∼ R/mP ∼
\ Y
(30) R/I0 = R/ = = k #Fq .
P ∈Fn
q P ∈Fn
q
222 11. NULLSTELLENSÄTZE

The Correspondence Theorem now gives us canonical bijections between the set of
ideals containing I0 and the set of subsets of Fnq . Since every subset of the finite set
n
Fnq is Zariski closed, there are precisely 2#Fq Zariski-closed subsets and therefore
n n
precisely 2#Fq ideals J with J = J. By (30) there are precisely 2#Fq ideals J
containing I0 , so we must have J = J for all such ideals. 
Remark: It seems that Theorem 11.19 was first stated and proved (as in the first
proof above) in a 1991 technical report of R. Germundsson [Ge91].

6. Terjanian’s Homogeneous p-Nullstellensatz


Theorem 11.20. Let k be an algebraically closed field, let 1 ≤ m < n be
integers, and let f1 , . . . , fm ∈ k[t1 , . . . , tn ]. Put
V := V (f1 , . . . , fm ) = {x = (x1 , . . . , xn ) ∈ k n | f1 (x) = . . . = fm (x) = 0}.
Then V is either empty or infinite.
Proof. Put I := hf1 , . . . , fm i. Seeking a contradiction, we suppose that V =
{p1 , . . . , pN } is nonempty and finite. For x ∈ k n we have x ∈ V iff I is contained
in the maximal ideal mx of polynomials that vanish at x, so the maximal ideals
contaning I are precisely mp1 , . . . , mpN . Since k[t1 , ..., tn ] is Jacobson, we have
N
\ N
Y
rad I = mpi = mpi .
i=1 i=1
QN
Let p be a minimal prime ideal over I. Then p ⊃ rad I = i=1 mpi , so p = mx for
some x = (x1 , . . . , xn ) ∈ k n . Applying the Generalized Principal Ideal Theorem to
p = mx we get that ht mxi ≤ m < n. But
(0) ( ht1 − x1 i ( ht1 − x1 , t2 − x2 i ( . . . ( ht1 − x1 , . . . , tn − xn i = mx
shows that mx has height at least n, a contradiction. 
Exercise 11.17. For which fields k does the conclusion of Theorem 11.20 hold?
Evidently not when k is finite!
a) Let f = t21 + t22 ∈ R[t1 , t2 ]. Show: V (f ) = {(0, 0)} is nonempty and finite.
b) Show: if k is not algebraically closed, there is f ∈ k[t1 , t2 ] such that
V (f ) = {(0, 0)}.
Corollary 11.21 (Homogeneous Nullstellensatz). Let k be an algebraically
closed field, m, n ∈ Z+ , and let f1 , . . . , fm ∈ k[t0 , . . . , tn ] be homogeneous polyno-
mials of positive degree. If m ≤ n, then there is 0 6= x ∈ k n+1 such that
f1 (x) = . . . = fn (x) = 0.
Proof. Let V be as in Theorem 11.20. Since each fi is homogeneous we have
0 ∈ V , and thus by Theorem 11.20 the set V is infinite. 
While Theorem 11.21 is a very classical result – it was used (not necessarily with
proper justification) by 19th century mathematicians studying varieties in projec-
tive space – the following generalization is a 1972 theorem of G. Terjanian.

Let p be a prime number. We say that a field k is a p-field if every every fi-
nite extension l/k has degree a power of p.
6. TERJANIAN’S HOMOGENEOUS p-NULLSTELLENSATZ 223

Examples: a) Every separably closed field is a p-field.


b) Every real-closed field is a 2-field.
c) A perfect field k is a p-field iff Gal(k/k) is a pro-p-group.
Theorem 11.22. (Terjanian’s Homogeneous p-Nullstellensatz) Let k be a p-
field, and let n ∈ Z+ . For 1 ≤ i ≤ n, let fi ∈ k[t0 , . . . , tn ] be homogeneous of degree
di indivisible by p. Then there is 0 6= x = (x1 , . . . , xn ) ∈ k n such that
f1 (x) = . . . = fn (x) = 0.
The proof given in [Te72] was rather involved; a significantly simpler proof is given
in [P], but even this involves more graded algebra than we wish to discuss here.
However, following Arason and Pfister [AP82] we will now deduce some striking
consequences of the Homogeneous 2-Nullstellensatz for real-closed fields.
Exercise 11.18. Let k be a field of characteristic different from 2, and let
f ∈ k[t1 , . . . , tn ]. We say that f is an odd polynomial if f (−t) = −f (t). Show:
an odd polynomial is a sum of monomials each of odd total degree.
Theorem 11.23. (Algebraic Borsuk-Ulam) Let k be a real closed field, n ∈ Z+ ,
and for 1 ≤ i ≤ n, let fi ∈ k[t1 , . . . , tn+1 ] be an odd polynomial: fi (−t) = −fi (t).
Then there is x = (x1 , . . . , xn+1 ) ∈ k n+1 such that
x21 + . . . + x2n+1 = 1, f1 (x) = . . . = fn (x) = 0.
Proof. So as to be able to apply Terjanian’s Homogeneous p-Nullstellensatz,
we homogenize: let t0 be an additional indeterminate and let f˜i ∈ k[t0 , . . . , tn+1 ]
be the unique homogeneous polynomial such that f˜i (1, t1 , . . . , tn+1 ) = fi , of degree
di = deg fi . Being an odd polynomial, each fi only contains monomials of odd
degree; thus each di is odd and t0 occurs in f˜i to even powers only. Thus we
may make the change of variables t20 7→ t21 + . . . + t2n+1 in each f˜i , leading to
homogeneous polynomials g1 , . . . , gn ∈ k[t1 , . . . , tn+1 ] of odd degrees d1 , . . . , dn .
Applying Theorem 11.22 with p = 2, we get 0 6= a ∈ k n+1 such that g1 (a) = . . . =
−1
gn (a) = 0. Since the gi ’s are homogeneous, we may scale by a21 + . . . + a2n+1 )
to get an a such that a21 + . . . + a2n+1 = 1 and g1 (a) = . . . = gn (a) = 0. Thus
fi (a) = f˜i (1, a) = gi (a) = 0 for all i, and we’re done. 
We now revert to the case of k = R. As usual, for x = (x1 , . . . , xn ) ∈ Rn , we put
q
||x|| = x21 + . . . + x2n };
for x, y ∈ Rn , we put
d(x, y) = ||x − y||,
and we define the unit sphere
S n = {x ∈ Rn | ||x|| = 1}
and the unit disk
Dn = {x ∈ Rn | ||x|| ≤ 1}.
A subset S ⊂ Rn is symmetric if x ∈ S =⇒ −x ∈ S. If S ⊂ Rm and T ⊂ Rn are
symmetric subsets, then f : S → T is odd if for all x ∈ S, f (−x) = −f (x). For
x ∈ S n , x and −x are antipodal, and {x, −x} is called an antipodal pair.
224 11. NULLSTELLENSÄTZE

Corollary 11.24. (Topological Borsuk-Ulam)


Let f : S n → Rn be a continuous, odd map. Then there is x ∈ S n with f (x) = 0.
Proof. Write f = (f1 , . . . , fn ), for fi : S n → R an odd continuous map.
Seeking a contradiction, we suppose f has no zero. Since S n is compact, there is
δ > 0 such that for all x ∈ S n , maxi |fi (x)| ≥ δ. Choose 0 <  < δ and apply the
Weierstrass Approximation Theorem to the continuous functions f1 , . . . , fn on the
compact space S n : there are p1 , . . . , pn ∈ R[t1 , . . . , tn+1 ] such that |fi (x)−pi (x)| < 
for all i and all x ∈ S n . Put qi (t) = 21 (pi (t) − pi (−t)); then for x ∈ S n ,
|fi (x) − fi (−x) − pi (x) + pi (−x)|
|fi (x) − qi (x)| =
2
|fi (x) − pi (x)| + |fi (−x) − pi (−x)|
≤ < .
2
It follows that for all x ∈ S n , maxi |qi (x)| ≥ δ −  > 0, so the qi ’s have no simulta-
neous zero on S n , contradicting Theorem 11.23. 

Corollary 11.25.
The following statements are equivalent – and hence, by Corollary 11.24, all true.
(i) Every continuous, odd map f : S n → Rn has a zero.
(ii) There is no continuous, odd map g : S n → S n−1 .
(iii) Every continuous map f : S n → Rn identifies an antipodal pair.
(iv) (Lusternik-Schnirelmann-Borsuk) Let {F1 , . . . , Fn+1 } be a covering family of
closed subsets of S n . Then some member of the family contains an antipodal pair.
Proof. (i) =⇒ (ii): Let ι : S n−1 ,→ Rn be the natural inclusion. If g : S n →
n−1
S is continuous and odd, ι ◦ g : S n → Rn is continuous and odd with no zero.
(ii) =⇒ (iii): If f identifies no antipodal pair, then g : S n → S n−1 by x 7→
f (x)−f (−x)
||f (x)−f (−x)|| is continuous and odd.
(iii) =⇒ (i): Let f : Rn → S n be odd. By assumption, there is x ∈ S n such that
f (x) = f (−x), but since also f (x) = −f (−x), we conclude f (x) = 0.
Sn+1
(iii) =⇒ (iv): Let F1 , . . . , Fn+1 be closed subsets of S n such that i=1 Fi = S n ;
suppose that none of the sets F1 , . . . , Fn contains an antipodal pair: equivalently,
putting Ei = −Fi , we have that Ei ∩ Fi = ∅ for 1 ≤ i ≤ n. For a point x and
a subset Y of S n , put d(x, Y ) = inf{d(x, y) | y ∈ Y }. For 1 ≤ i ≤ n + 1, define
fi : S n → R by fi (x) = d(x, Ei ) − d(x, Fi ). Observe that
x ∈ Fi =⇒ fi (−x) < 0 < fi (x),
x ∈ Ei =⇒ fi (x) < 0 < fi (−x).
Applying condition (iii) to f = (f1 , . . . , fn ) : S n → Rn , we get x0 ∈ S n such that
f (−x0 ) = f (x0 ). Thus neither x0 nor −x0 lies in any Fi with 1 ≤ i ≤ n, hence
both x0 and −x0 must lie in Fn+1 .
(iv) =⇒ (ii): Let f : S n → S n−1 be continuous. Observe the following “converse”
to Lusternik-Schnirelmann-Borsuk: there is a covering family {Ei }n+1 i=1 of closed
subsets of S n−1 , each of diameter less than 2. (We leave the verification of this as
an exercise.) For 1 ≤ i ≤ n+1, put Fi = f −1 (Ei ). Thus {Fi }n+1 i=1 is a covering of S
n
n
by closed subsets, so by condition (iv) for some 1 ≤ i ≤ n + 1 and x0 ∈ S we have
x0 , −x0 ∈ Fi , i.e., f (x0 ), f (−x0 ) ∈ Ei . Since Ei has diameter less than 2, it contains
no antipodal pair, and thus f cannot be odd, for otherwirse f (x0 ), −f (x0 ) ∈ Ei . 
6. TERJANIAN’S HOMOGENEOUS p-NULLSTELLENSATZ 225

Exercise 11.19. Verify: for any n ∈ Z+ , S n can be covered by n + 2 closed


subsets each of diameter less than 2.
(Suggestion: take a regular simplex inscribed in Dn and consider the projections of
its faces onto S n .)
For a function f : Rn → R, let us put
f + = {x ∈ Rn | f (x) > 0},

f 0 = {x ∈ Rn | f (x) = 0},
f − = {x ∈ Rn | f (x) < 0}.
For a Lebesgue measurable subset S ⊂ Rn , we denote its measure by Vol(S).

The following result is due to Stone and Tukey [ST42].


+
Corollary
n+d
 11.26. (Polynomial Ham nSandwich Theorem) Let d, n ∈ Z and
put N = d − 1. Let U1 , . . . , UN ∈ R be measurable, finite volume subsets.
There is a polynomial P ∈ R[t1 , . . . , tn ] of degree at most d which bisects each Ui :
∀1 ≤ i ≤ N, Vol(Ui ∩ P + ) = Vol(Ui ∩ P − ).
Proof. Let Vd ⊂ R[t1 , . . . , tn ] be the R-subspace of polynomials of total degree
at most d, so dim Vd = N + 1. Endow Vd with a norm || · || (all norms on a finite-
dimensional real vector space are equivalent, so we need not be more specific than
this). Let S N be the unit sphere in Vd . We define a function
f = (f1 , . . . , fN ) : S N → RN
by
fi (P ) = Vol(Ui ∩ P + ) − Vol(Ui ∩ P − ).
Step 1: We claim that f is continous.
proof of claim: One easily reduces to the claim that for any measurable, finite
volume subset U ⊂ Rn , the mapping
M : P ∈ Vd• 7→ Vol(U ∩ P + )
is continuous. For this, let {Pn } be a sequence in Vd such that Pn → P with
respect to || · ||. It follows that Pn → P pointwise on (Rn hence in particular on)
U . Since Vol(U ) < ∞, we may apply Egorov’s Theorem: for each  > 0, there is
a measurable subset E ⊂ U with Vol(E) <  and such that Pn → P uniformly on
U \ E. Since Vol(P 0 ) = 0 and Vol(U ) < ∞, there is δ > 0 such that
Vol({x ∈ U | |P (x)| < δ}) < .
+
Take N ∈ Z such that for all n ≥ N , |Pn (x) − P (x)| < δ for all x ∈ U \ E. Then
| Vol(U ∩ Pn+ ) − Vol(U ∩ P + )| < 2.
Step 2: It is immediate that f is odd. By Corollary 11.24, there is P ∈ S N such
that f (P ) = 0, and such a P bisects each Ui . 

Corollary 11.27. (No Retraction Theorem) There is no retraction from Dn


to S n−1 , i.e., no continuous map r : Dn → S n−1 such that r|S n−1 = 1S n−1 .
226 11. NULLSTELLENSÄTZE

Proof. Let π : Rn+1 → Rn , (x1 , . . . , xn , xn+1 ) 7→ (x1 , . . . , xn ), and let


Hn+ = {(x1 , . . . , xn+1 ) ∈ S n | xn+1 ≥ 0}, Hn− = {(x1 , . . . , xn+1 ) ∈ S n | xn+1 ≤ 0}.
Suppose r : Dn → S n−1 is a retraction, and define g : S n → S n−1 by
r(−π(x)), x ∈ Hn+ ,

g(x) =
−r(π(x)), x ∈ Hn−
Then g is well-defined, continuous and odd, contradicting Corollary 11.25b). 
Corollary 11.28. (Brouwer Fixed Point Theorem) Each continuous function
f : Dn → Dn has a fixed point.
Proof. Suppose f : Dn → Dn is continuous with f (x) 6= x for all x ∈ Dn . For
x ∈ Dn , consider the ray rx with initial point f (x) and lying on the line determined
by f (x) and x. Then rx intersects S n−1 at a unique point, say r(x), and x 7→ r(x)
defines a retraction r : Dn → S n−1 , contradicting Corollary 11.27. 
CHAPTER 12

Goldman domains and Hilbert-Jacobson rings

1. Goldman domains
Lemma 12.1.
For a domain R with fraction field K, the following are equivalent:
(i) K is finitely generated as an R-algebra.
(ii) There exists f ∈ K such that K = R[f ].
Proof. Of course (ii) =⇒ (i). Conversely, if K = R[f1 , . . . , fn ], then write
fi = pqii , and then K = R[ q1 ···q
1
n
]. 
A ring satisfying the conditions of Lemma 12.1 will be called a Goldman domain.
Exercise 12.1. Show: an overring1 of a Goldman domain is a Goldman do-
main.
Lemma 12.2. Let R be a domain with fraction field K, and 0 6= x ∈ R. the
following are equivalent:
(i) Any nonzero prime ideal of R contains x.
(ii) Any nonzero ideal contains a power of x.
(iii) K = R[x−1 ].
Proof. (i) =⇒ (ii): let I be a nonzero ideal. If I is disjoint from {xn }, then
by Multiplicative Avoidance (5.24), I can be extended to a prime ideal disjoint
from {xn }, contradicting (i).
(ii) =⇒ (iii): Let 0 6= y ∈ R. By (ii), we have (y) contains some power of x,
say xk = yz. But this implies that y is a unit in R[x−1 ].
(iii) =⇒ (i): The prime ideals killed in the localization map R 7→ R[x−1 ] are
precisely those which meet the multiplicatively closed set {xk }, i.e., contain x. 
Corollary 12.3. For a domain R, the following are equivalent:
(i) R is a Goldman domain.
(ii) The intersection of all nonzero prime ideals of R is nonzero.
Exercise 12.2. Prove Corollary 12.3.
Easy examples of Goldman domains: a field, k[[t]], Z(p) . In fact we have devel-
oped enough technology to give a remarkably clean characterization of Noetherian
Goldman domains.
Theorem 12.4. Let R be a domain.
a) If R has only finitely many primes, then R is a Goldman domain.
b) If R is a Noetherian Goldman domain, then R has finitely many primes.
c) A Noetherian Goldman domain is either a field or a one-dimensional domain.
1An overring of a domain R is a ring intermediate between R and its fraction field K.

227
228 12. GOLDMAN DOMAINS AND HILBERT-JACOBSON RINGS

Proof. Throughout we may – and shall – assume that R is not a field.


a) Suppose that R has only finitely many primes, and let p1 , . . . , pn be the
nonzero prime ideals of R. For 1 ≤ i ≤ n, let 0 6= xi ∈ pi , and put x = x1 · · · xn .
Then the multiplicative set S generated by x meets every nonzero prime of R, so
that S −1 R has only the zero ideal. In other words, R[ x1 ] is the fraction field of R,
so R is a Goldman domain. (Alternately, this follows quickly from Corollary 12.3.)
b) Similarly, for a Goldman domain R we can write K = R[ x1 ] for x ∈ R and
then every nonzero prime of R contains x. Suppose first that (x) itself is prime,
necessarily of height one by the Hauptidealsatz (Theorem 8.44), hence if R has any
primes other than (0) and (x) – especially, if it has infinitely many primes – then it
has a height two prime q. But by Corollary 8.48 a Noetherian ring cannot have a
height two prime unless it has infinitely many height one primes, a contradiction.
So we may assume that (x) is not prime, and then the minimal primes of the
Noetherian ring R/(x) are finite in number – say p1 , . . . , pn – and correspond to
the primes of R which are minimal over x, so again by the Hauptidealsatz they all
have height one. Similarly, if R has infinitely many primes there would be, for at
least one i (say i = 1), a height two prime q ⊃ p1 . But then by Corollary 8.48
the “interval” (0, q) is infinite. Each element of this set is a height one prime ideal
containing (x), i.e., is one of the pi ’s, a contradiction. Part c) follows by again
applying Corollary 8.48: a Noetherian ring of dimension at least two must have
infinitely many primes. 
Remark: a non-Noetherian Goldman domain can have infinitely many primes
and/or primes of arbitrarily large height.
Proposition 12.5. Let R be a domain. Then the polynomial ring R[t] is not
a Goldman domain.
Proof. Let K be the fraction field of R. If R[t] is a Goldman domain, then
by Exercise 12.1, so is K[t]. But K[t] is a Noetherian domain with infinitely many
primes – e.g., Euclid’s proof of the infinitude of primes in Z carries over verbatim
to K[t] – so Theorem 12.4 applies to show that K[t] is not a Goldman domain. 
Proposition 12.6. Let R be a domain, and T ⊃ R an extension domain which
is algebraic and finitely generated as an R-algebra. Then R is a Goldman domain
iff T is a Goldman domain.
Proof. Let K and L be the fraction fields of R and T , respectively. Suppose
first that R is a Goldman domain: say K = R[ u1 ]. Then T [ u1 ] is algebraic over
the field K, so is a field, hence we have L = T [ u1 ]. Conversely, suppose that T
is a Goldman domain: say L = T [ v1 ]; also write T = R[x1 , . . . , xk ]. The elements
v −1 , x1 , . . . , xk are algebraic over R hence satisfy polynomial equations with coeffi-
cients in R. Let a be the leading coefficient of a polynomial equation for v −1 and
b1 , . . . , bk be the leading coefficients of polynomial equations for x1 , . . . , xk . Let
R1 := R[a−1 , b−1 −1
1 , . . . , bk ]. Now L is generated over R1 by x1 , . . . , xk , v
−1
, all of
which are integral over R1 , so L is integral over R1 . Since L is a field, it follows that
R1 is a field, necessarily equal to K, and this shows R is a Goldman domain. 
Corollary 12.7. Let R ⊂ S be an inclusion of domains, with R a Goldman
domain. Suppose that u ∈ S is such that R[u] is a Goldman domain. Then u is
algebraic over R, and R is a Goldman domain.
1. GOLDMAN DOMAINS 229

Theorem 12.8. For a domain R, the following are equivalent:


(i) R is a Goldman domain.
(ii) There exists a maximal ideal m of R[t] such that m ∩ R = (0).
Proof. (i) =⇒ (ii): We may assume WLOG that R is not a field. Write
K = R[ u1 ]. Define a homomorphism ϕ : R[t] → K by sending t 7→ u1 . Evidently ϕ
is surjective, so its kernel m is a maximal ideal, and clearly we have m ∩ R = 0.
(ii) =⇒ (i): Suppose m is a maximal ideal of R[t] such that m ∩ R = (0). Let
v be the image of t under the natural homomorphism R[t] → R[t]/m. Then R[v] is
a field, so by Corollary 12.7, R is a Goldman domain. 
We define a prime ideal p of a ring R to be a Goldman ideal if R/p is a Goldman
domain. Write G Spec R for the set of all Goldman ideals. Thus a Goldman ideal
is more general than a maximal ideal but much more special than a prime ideal.
Proposition 12.9. Let R be a ring and I an ideal of R.
a) The nilradical of R is the intersection of all Goldman ideals of R.
b) The radical of I is the intersection of all Goldman ideals containing I.
T T
Proof. a) We know that N = p∈Spec R p, so certainly N ⊂ p∈G Spec R p.
Conversely, suppose x ∈ R\N . The ideal (0) is then disjoint from the multiplicative
set S = {xn }. By multiplicative avoidance, we can extend (0) to an ideal p maximal
with respect to disjointness from S. We showed earlier that p is prime; we now claim
that it is a Goldman ideal. Indeed, let x denote the image of x in R = R/p. By
maximality of p, every nonzero prime of R contains x. By Lemma 12.2, this implies
R[x−1 ] is a field, thus R is a Goldman domain, and therefore p is a Goldman ideal
which does not contain x. Part b) follows by correspondence, as usual. 
The following result may seem completely abstruse at the moment, but soon enough
it will turn out to be the key:
Corollary 12.10. An ideal I in a ring R is a Goldman ideal iff it is the
contraction of a maximal ideal in the polynomial ring R[t].
Proof. This follows from Theorem 12.8 by applying the correspondence prin-
ciple to the quotient ring R/I. 
Theorem 12.11. a) Let M be a maximal ideal in R[t], and suppose that its
contraction m = M ∩ R is maximal in R. Then M can be generated by m and by
one additional element f , which can be taken to be a monic polynomial which maps
modulo m to an irreducible polynomial in R/m[t].
b) If, moreover, we suppose that R/m is algebraically closed, then M = hm, t − ai
for some a ∈ R.
Proof. a) Since M contains m, by correspondence M may be viewed as a
maximal ideal of R[t]/mR[t] ∼ = (R/m)[t], a PID, so corresponds to an irreducible
polynomial f ∈ R/m[t]. If f is any lift of f to R[t], then M = hm, f i. Part b)
follows immediately from the observation that an irreducible univariate polynomial
over an algebraically closed field is linear. 
The following more elementary result covers the other extreme.
Theorem 12.12. Let R be a domain, with fraction field K. Let ι : R[t] → K[t]
be the natural inclusion. Then ι∗ induces a bijection between the prime ideals P of
R[t] such that P ∩ R = {0} and the prime ideals of K[t].
230 12. GOLDMAN DOMAINS AND HILBERT-JACOBSON RINGS

Proof. Let S = R \ {0}. The key observation is that S −1 R[t] = K[t]. Recall
(Proposition 7.5) that in any localization map R 7→ S −1 R, the prime ideals which
push forward to the unit ideal are precisely those which meet S, whereas the local-
ization map restricted to all other prime ideals is a bijection onto the set of prime
ideals of S −1 R. Applying that in this case gives the desired result immediately! 

2. Hilbert rings
To put Theorem 12.11 to good use, we need to have a class of rings for which the
contraction of a maximal ideal from a polynomial ring is again a maximal ideal. It
turns out that the following is the right class of rings:

Definition: A Hilbert ring is a ring in which every Goldman ideal is maximal.


Proposition 12.13. Any quotient ring of a Hilbert ring is a Hilbert ring.
Proof. This follows immediately from the correspondence between ideals of
R/I and ideals of R containing I. 
A direct consequence of the definition and Proposition 12.9 is the following:

T Proposition 12.14. Let I be an ideal in a Hilbert ring R. Then the intersection


m sup I m of all maximal ideals m containing I is rad(I).

Examples: Any zero dimensional ring is a Hilbert ring. Especially, a field is a


Hilbert ring, as is any Artinian ring or any Boolean ring.
Exercise 12.3. a) Let R be a one-dimensional Noetherian domain. the fol-
lowing are equivalent:
(i) R is a Hilbert ring.
(ii) The Jacobson radical of R is 0.
(iii) R has infinitely many prime ideals.
(iv) R is not a Goldman domain.
b) Deduce: the ring Z of integers is a Hilbert domain.
Theorem 12.15. Let R be a Hilbert ring, and S a finitely generated R-algebra.
Then:
a) S is also a Hilbert ring.
b) For every maximal ideal P of S, p := P ∩ R is a maximal ideal of R.
c) The degree [S/P : R/p] is finite.
Proof. a) It suffices to show that R is a Hilbert ring iff R[t] is a Hilbert ring,
for then, if R is a Hilbert ring, by induction any polynomial ring R[t1 , . . . , tn ] is a
Hilbert ring, and any finitely generated R-algebra is a quotient of R[t1 , . . . , tn ] and
thus a Hilbert ring. Note also that since R is a homomorphic image of R[t], if R[t]
is a Hilbert domain, then so also is R.
So suppose R is a Hilbert ring, and let q be a Goldman ideal in R[t]; we must
show q is maximal. Put p = q ∩ R. As above, we can reduce to the case p = 0, so
in particular R is a domain. Let a be the image of t in the natural homomorphism
R[t] → R[t]/q. Then R[a] is a Goldman domain. By Corollary 12.7, a is algebraic
over R, and R is a Goldman domain. But since we assumed that R was a Hilbert
ring, this means that R is a field, and thus R[a] = R[t]/q is a field, so q is maximal.
b) We may write S = R[t1 , . . . , tn ]/I. A maximal ideal m of S is just a maximal
3. JACOBSON RINGS 231

ideal of R[t1 , . . . , tn ] containing I. By Corollary 12.10, the contraction m0 of m to


R[t1 , . . . , tn−1 ] is a Goldman ideal of the Hilbert ring R[t1 , . . . , tn−1 ], so is therefore
maximal. Moreover, by Theorem 12.11, m is generated by m0 and an irreducible
polynomial in R/m0 [t], so that the residual extension R[t1 , . . . , tn ]/m has finite
degree over R[t1 , . . . , tn−1 /m0 . Again, induction gives the full result. 
Applying Theorem 12.15c) in the case R = k is a field, we deduce our second proof
of Zariski’s Lemma (Lemma 11.1).
Theorem 12.16. Let R be a Noetherian Hilbert ring. Then
dim(R[t]) = dim R + 1.
Proof. Let 0 = p0 ( p1 ( . . . ( pd be a chain of prime ideals in R. Then,
with ι : R ,→ R[t] the natural inclusion,
ι∗ p0 ( . . . ( ι∗ pn ( hι∗ (pn ), ti
is a chain of prime ideals of R[t] of length d + 1, hence for any ring R we have
dim R[t] ≥ dim R+1. Conversely, it suffices to show that the height of any maximal
ideal P of R[t] is at most d + 1. For this, put p = P ∩ R. By Theorem 12.15, p
is maximal in R, so Theorem 12.11 tells us that there exists f ∈ R[t] such that
P = hι∗ p, f i. Applying Krull’s Hauptidealsatz (Theorem 8.44) in the quotient ring
R[t]/ι∗ p, we get that the height of P is at most one more than the height of p. 
Corollary 12.17. Let k be a field, and put R = k[t1 , . . . , tn ].
a) Then every maximal ideal of R has height n and can be generated by n elements
(and no fewer, by Theorem 8.49).
b) In particular, dim R = n.
Exercise 12.4. Prove Corollary 12.17.

3. Jacobson Rings
Theorem 12.18. For a ring R, the following are equivalent:
(i) For all I ∈ I(R), r(I) is the intersection of all maximal ideals containing I.
(i0 ) In every quotient ring of R, the nilradical equals the Jacobson radical.
(ii) Every prime ideal p of R is the intersection of all maximal ideals containing p.
(iii) Every nonmaximal prime ideal p of R is equal to the intersection of all prime
ideals strictly containing p.
If R satisfies these equivalent properties it is called a Jacobson ring.
Proof. (i) ⇐⇒ (i0 ) is immediate from the Correspondence Theorem. T
(i) =⇒ (ii): If (i) holds, then in particular for any radical ideal I, I = m⊃I m,
and prime ideals are radical.
(ii) =⇒ (i): for any ideal I of R,
\ \ \ \
rad I = p= m= m.
p⊃I p⊃I m⊃p m⊃I
T
(ii) =⇒ (iii): If p is prime but not maximal, then p = m⊃p m and all the maximal
ideals containing p strictly contain p.
¬ (ii) =⇒ ¬ (iii): Let p be a prime which is not the intersection of the maximal
ideals containing it. Replacing R with R/p, we may assume R is a domain with
nonzero Jacobson radical J(R). Let x ∈ J(R) \ {0}, and choose, by Multiplicative
232 12. GOLDMAN DOMAINS AND HILBERT-JACOBSON RINGS

Avoidance, an ideal p which is maximal with respect to the property that x ∈ / p.


Since x ∈/ J(R) \ p, p is not maximal; since x lies in every ideal properly containing
p, p is not equal to the intersection of prime ideals strictly containing it. 

Corollary 12.19. Every quotient ring of a Jacobson ring is Jacobson.


Proof. This is immediate from condition (i0 ) of Theorem 12.18. 

4. Hilbert-Jacobson Rings
Proposition 12.20. Suppose R is both a Goldman domain and a Jacobson
ring. Then R is a field.
Proof. Let K be the fraction field of R, and suppose for a conradiction that
R 6= K. Then there is a nonzero nonunit f ∈ R such that K is the localization
of R at the multiplicative subset S = {f, f 2 , . . .}. Let m be a maximal ideal of R.
Since R is not a field, m is not zero, and thus the pushforward of R to S −1 R is the
unit ideal. By Proposition 7.5, m meets S. Since m is prime, we conclude f ∈ m.
It follows that the Jacobson radical of R contains f is accordingly nonzero. On the
other hand R, being a domain, has zero nilradical. Thus R is not Jacobson. 

Theorem 12.21. For a commutative ring R, the following are equivalent:


(i) R is a Hilbert ring.
(ii) R is a Jacobson ring.
(iii) For all maximal ideals m of R[t], m ∩ R is a maximal ideal of R.
(iv) (Zariski’s Lemma) Let K be a field that is finitely generated as an R-algebra.
Then K is finitely generated as a R-module.
Proof. (i) =⇒ (ii) by Proposition 12.14.
(ii) =⇒ (i): Suppose R is Jacobson and p is a Goldman ideal of R. Then R/p
is a Goldman domain (by definition of Goldman ideal) and a Jacobson ring (by
Corollary 12.19), hence a field (by Proposition 12.20), so p is maximal.
(ii) =⇒ (iii) is Theorem 12.15b).
(iii) =⇒ (i): Suppose R is a ring such that every maximal ideal of R[t] contracts
to a maximal ideal of R, and let p be a Goldman ideal of R. By Corollary 12.10, p
is the contraction of a maximal ideal of R[t], hence by assumption p is maximal.
(i) =⇒ (iv) by Theorem 12.15c).
(iv) =⇒ (ii): By Theorem 12.18, it suffices to show that every nonmaximal prime
p is the intersection of the prime ideals strictly containing it. That is, let x ∈ R \ p:
we will find a prime ideal q ) p such that x ∈ / q. Let B be the domain R/p, so the
image of x in B (which we continue to denote by x) is nonzero. Then B 0 = B[ x1 ]
is a finitely generated R-algebra. If B 0 is a field, then by hypothesis B 0 is finitely
generated as an R-module and thus, equivalently, finitely generated as a B-module.
But this implies that B is a field, a basic fact about integral extensions which
will be proved later on in the notes (Theorem 14.1, Propostion 14.8a)) and thus
p is maximal, contradiction. So B 0 is not a field and thus it contains a nonzero
maximal ideal, whose pullback to B is a prime ideal q not containing x. The ideal
q corresponds to a prime ideal q ) p of R not containing x. 

In the sequel we will use the consolidated terminology Hilbert-Jacobson ring for
a ring satisfying the equivalent conditions of Theorem 12.21.
5. APPLICATION: ZERO-DIMENSIONAL IDEALS IN POLYNOMIAL RINGS 233

5. Application: Zero-Dimensional Ideals in Polynomial Rings


Let k be a field with algebraic closure k, and let I be an ideal of the polynomial
ring k[t1 , . . . , tn ]. Recall (from §11.2.1) that V a (I) denotes the set of simultaneous
n
zeros of I in k .

Theorem 12.22. For an ideal I of k[t1 , . . . , tn ], the following are equivalent:


(i) dimk k[t1 , . . . , tn ]/I is finite.
(ii) The ring k[t1 , . . . , tn ]/I is Artinian.
(iii) The ring k[t1 , . . . , tn ]/I has Krull dimension 0.
(iv) The ring k[t1 , . . . , tn ]/I has only finitely many prime ideals.
(v) The ring k[t1 , . . . , tn ]/I has only finitely many maximal ideals.
(vi) The set V a (I) is finite.
An ideal satisfying these conditions is called zero-dimensional.

Proof. (i) =⇒ (ii): A finite dimensional k-algebra A has no infinite descend-


ing chains of k-submodules, let alone A-submodules.
(ii) ⇐⇒ (iii): By the Hilbert Basis Theorem the ring k[t1 , . . . , tn ] is Noetherian
hence so is its quotient k[t1 , . . . , tn ]. A ring is Artinian iff it is Noetherian of Krull
dimension zero.
(ii) =⇒ (iv): An Artinian ring has finitely many prime ideals, all of which are
maximal.
(iv) =⇒ (v) is immediate.
(v) =⇒ (vi): We have V a (I) = V a (rad I). Since k[t1 , . . . , tn ] is a Hilbert-Jacobson
ring, rad I is the intersection of all maximal ideals containing I, so I satisfies con-
dition (v) iff rad I does and we may assume that I is a radical ideal. Because there
TN
are finitely many maximal ideals m1 , . . . , mN , we have I = i=1 mi , so
n
[
V a (I) = V a (mi ),
i=1
a
so it suffices to show that V (m) is finite for a maximal ideal m. By Zariski’s
Lemma, k[t1 , . . . , tn ]/m is a finite-degree field extension of k, so there are only
n
finitely many k-algebra homomorphisms k[t1 , . . . , tn ]/m ,→ k . However, if x =
(x1 , . . . , xn ) ∈ V a (m) then the k-algebra homomorphism Ex : k[t1 , . . . , tn ] → k that
maps each ti to xi has m in its kernel, hence induces a k-algebra homomorphism
Ex : k[t1 , . . . , tn ]/m → k, and the association x ∈ V a (m) 7→ Ex is injective since
x = (Ex (t1 ), . . . , Ex (tn )). Thus V a (m) is finite.
(vi) =⇒ (i): Let I = Ik[t1 , . . . , tn ]. Then V a (I) = V (I) and

k[t1 , . . .n ]/I = k[t1 , . . . , tn ]/I ⊗k k

and thus dimk k[t1 , . . . , tn ]/I is finite iff dimk k[t1 , . . . , tn ]/I is finite. So we may
assume that k is algebraically closed, and thus the points x of V (I) correspond
precisely to the maximal ideals of k[t1 , . . . , tn ] containing I, so we are assuming
that I is contained in only finitely many maximal ideals, say m1 , . . . , mN . Thus

k[t1 , . . . , tn ]/ rad I ∼
= k[t1 , . . . , tn ]/m1 × . . . × k[t1 , . . . , tn ]/mN ,
so dimk k[t1 , . . . , tn ]/ rad I is finite by Zariski’s Lemma. In particular k[t1 , . . . , tn ]/ rad I
has Krull dimension zero, hence so does k[t1 , . . . , tn ]/I, hence it is Artinian. By
234 12. GOLDMAN DOMAINS AND HILBERT-JACOBSON RINGS

Theorem 8.37, for all sufficiently large a we have


n
k[t1 , . . . , tn ]/I ∼
Y
= k[t1 , . . . , tn ]/mai .
i=1

For each 0 ≤ j ≤ a − 1, mji /mj+1


is finitely generated over k[t1 , . . . , tn ]/mi , hence
i
finite-dimensional over k, and thus dimk k[t1 , . . . , tn ]/I is finite. 
Exercise 12.5. Let I be an ideal of k[t1 , . . . , tn ].
a) Show: I is zero-dimensional iff rad I is zero-dimensional.
n
b) Show: if S ⊂ k is finite, then I(S) is a radical zero-dimensional ideal.
c) Show: every radical zero-dimensional ideal of k[t1 , . . . , tn ] is of the form I(S) iff
k is algebraically closed.
Exercise 12.6. a) Let I be an ideal of k[t1 , . . . , tn ]. Show: I is zero-dimensional
radical iff k[t1 , . . . , tn ]/I is a finite product of fields (equivalently, is semisimple).
b) Let I ⊂ J be ideals of k[t1 , . . . , tn ]. Show: if I is zero-dimensional, so is J. If I
is zero-dimensional radical, so is J.
According to Corollary 12.17, every maximal ideal of k[t1 , . . . , tn ] can be generated
by n elements. In general, if I1 , . . . , IN are ideals and for each 1 ≤ j ≤ N , Sj is a
finite set of generators of Ij , then {f1 · · · fn | fj ∈ Sj } is a finite set of generators of
I1 · · · IN . Thus a radical zero-dimensional ideal I of k[t1 , . . . , tn ] that is contained
in N maximal ideals has a generating set of cardinality nN . But actually we can
do much better than this.
Theorem 12.23. Let I be a radical zero-dimensional ideal of R = k[t1 , . . . , tn ].
Then I can be generated by n elements.
TN
Proof. (Vasconcelos) Let I = i=1 mi with mi maximal ideals of k[t1 , . . . , tn ].
TN
Consider the ideal I ∩ k[t1 ] = I=1 mi ∩ k[t1 ]. By Theorem 12.15, each mi ∩ k[t1 ]
is a maximal ideal of k[t1 ], so I ∩ k[t1 ] is a radical ideal of the PID k[t1 ], so it is
generated by a product of nonassociate irreducible polynomials, say
f (t1 ) = g1 (t1 ) · · · gs (t1 ).
For 1 ≤ i ≤ s, let li = k[t1 ]/gi , a finite degree field extension of k. By CRT, we
have
s
R/(f R) = k[t1 ]/(f )[t2 , . . . , tn ] ∼
Y
= li [t2 , . . . , tn ].
i=1
For 1 ≤ i ≤ s, let Ii := IR/gi R = Ili [t2 , . . . , tn ]. Since (R/gi R)/Ii = R/hI, gi i,
by Exercise 12.6 each Ii is a radical zero-dimensional ideal of li [t2 , . . . , tn ], so by
induction there are fi,1 , . . . , fi,n−1 ∈ I such that hfi,1 , . . . , fi,n−1 iR/gi R = Ii .
For 1 ≤ i ≤ n, put Y
Gi (t1 ) := gj (t1 ),
j6=i
and for 1 ≤ k ≤ n − 1 put
X
Hk = Gj (t1 )fj,k .
1≤j≤s

We claim that
I = hf (t1 ), H1 , . . . , Hn−1 i.
5. APPLICATION: ZERO-DIMENSIONAL IDEALS IN POLYNOMIAL RINGS 235

Certainly we have
hf (t1 )i ⊂ hf (t1 ), G1 , . . . , Gn−1 i ⊂ I,
so passing to R/(f (t1 )) it is enough to show equality after localizating at each prime
ideal p containing f (t1 ). Let p be such a prime ideal. Then p contains gi (t1 ) for
some j, and since the gj ’s are pairwise comaximal, for all j 6= i, we have that gj (t1 )
is a unit modulo p, hence Gi (t1 ) is a unit modulo p, whereas Gj (t1 ) ∈ (gi ) for all
j 6= i. Therefore
hf (t1 ), H1 , . . . , Hn−1 ip = hgi (t1 ), fi,1 , . . . , fi,n−1 ip .
Since fi,1 , . . . , fi,n−1 generate I modulo (gi ), they generate I modulo p and thus Ip
modulo Rp , so by Nakayama’s Lemma they generate Ip , completing the proof. 
In particular, if S ⊂ k n is any finite subset, there are polynomials f1 , . . . , fn ∈
k[t1 , . . . , tn ] such that
{x ∈ k n | f1 (x) = . . . = fn (x) = 0} = S.
Exercise 12.7. Let I = ht1 , t2 i2 = ht21 , t1 t2 , t22 i in k[t1 , t2 ]. Show: I is zero-
dimensional and cannot be generated by 2 elements. (Thus the word “radical” in
the statement of Theorem 12.23 is essential.)
CHAPTER 13

Spec R as a topological space

1. The Prime Spectrum


For a ring R, we denote the set of all prime ideals of R by Spec R. Moreover, we
refer to Spec R as the Zariski spectrum – or prime spectrum – of R.

It is important to notice that Spec R comes with additional structure. First, it


has a natural partial ordering, in which the maximal elements are the maximal
ideals, and the minimal elements are (by definition) the minimal primes. Also,
as O. Zariski first observed, Spec R can be endowed with a topology. To see this,
for any ideal I of R, put V (I) = {p ∈ Spec R | p ⊃ I}.
Proposition 13.1. For ideals I, J of R, we have V (I) = V (J) iff rad I =
rad J.
Proof. For any ideal I and any prime ideal p, p ⊃ I iff p ⊃ rad I, and
therefore V (I) = V (rad I). Conversely, if V (I) = V (J), then the set of prime
ideals containing I is the same as the set of prime ideals containing J. So
\ \
rad I = p= p = rad J. 
p⊃I p⊃J

Exercise 13.1. Show: we have V (I) ⊂ V (J) iff rad I ⊃ rad J.


Now we claim that the family of subsets V (I) of Spec R has the following properties:

(ZT1) ∅ = V (R), Spec R = V ((0)). T


(ZT2) If {Ii } is any collection of ideals ofSR, then i V (Ii ) = V (hIi i). T
n n
(ZT3) If I1 , . . . , In are ideals of R, then i=1 V (Ii ) = V (I1 · · · In ) = V ( i=1 Ii ).
T
(ZT1) is obvious. As for (ZT2), let p be a prime ideal of R. Then p ∈ i V (Ii ) for
all i iff p ⊃ Ii for all i iff p contains the ideal generated by all Ii . As for (ZT3), p
contains a product of ideals iff it contains one of the ideals of the product.

Therefore there is a unique topology on Spec R in which the closed sets are precisely
those of the form V (I). This is called the Zariski topology.

It is of course natural to ask for a characterization of the open sets. Recall that a
base for the open sets of a topology is a collection {Bi } of open sets such that:

(BT1) for any point x ∈ Bi ∩ Bj , there exists a k such that x ∈ Bk ⊂ Bi ∩ Bj ;


(BT2) every open set is a union of the Bi ’s contained in it.
237
238 13. Spec R AS A TOPOLOGICAL SPACE

For f ∈ R, we define U (f ) := Spec R \ V ((f )). In other words, U (f ) is the


collection of all prime ideals which do not contain the element f . For f, g ∈ R,
U (f ) ∩ U (g) is the set of prime ideals p containing neither f nor g; since p is prime,
this is equivalent to p not containing f g, thus
U (f ) ∩ U (g) = U (f g),
which is a stronger property than (BT1). Moreover, any open set U is of T the form
Spec R\V (I). Each ideal I is the union of all of its elements fi , so V (I) = i V (fi ),
so that
\ [ [
U = Spec R \ V (I) = Spec R \ V (fi ) = (Spec R \ V (fi )) = U (fi ).
i i i

Proposition 13.2. Let R be any ring, and consider the canonical homomor-
phism f : R → Rred = R/ nil(R). Then f −1 : Spec Rred → Spec R is a homeomor-
phism.
Exercise 13.2. Prove Proposition 13.2.
Exercise 13.3. Let R1 , . . . , Rn be finitely many rings.`Show: Spec(R1 × . . . ×
n
Rn ) is canonically homeomorphic to the topological space i=1 Spec Ri .
Exercise 13.4. Let R be a Boolean ring. Earlier we defined a topology on the
set “M (R)” of all maximal ideals of R. But, as we know, a Boolean ring all prime
ideals are maximal, so as sets M (R) = Spec R. Show that moreover the topology
we defined on M (R) is the Zariski topology on Spec R.

2. Properties of the spectrum: quasi-compactness


More than sixty years ago now, N. Bourbaki introduced the term quasi-compact
for a topological space X for which any open covering has a finite subcovering.
The point of this terminology is to reserve compact for a space which is both
quasi-compact and Hausdorff, and thus emphasize that most of the nice properties
of compact spaces in classical topology do rely on the Hausdorff axiom. Nowhere
is this terminology more appropriate than in the class of spectral spaces, which
as we have seen above, are only Hausdorff in the comparatively trivial case of a
zero-dimensional ring. On the other hand:
Proposition 13.3. For any commutative ring R, Spec R is quasi-compact.
Proof. Let {Ui } be any open covering of Spec R. For each p ∈ Spec R, there
exists an element U of the cover containing p, and thus a principal open set X(f )
containing p and contained in U . Therefore there is a refinement of the cover
consisting of principal open subsets, and if this refinement has a finite cover, then
the original cover certainly does as well. Thus it suffices to assume that the Ui ’s
are basic open sets.1 So now suppose that Spec R = i U (fi ). Then we have
S
[ [ \
Spec R = U (fi ) = (Spec R \ V (fi )) = Spec R \ V (fi ),
i i i

1This is just the familiar, and easy, fact that it suffices to verify quasi-compactness on any
base for the topology. It is also true, but deeper, that one can verify quasi-compactness on any
subbase: Alexander’s Subbase Theorem.
4. PROPERTIES OF THE SPECTRUM: SEPARATION AND SPECIALIZATION 239

T
so that ∅ = i V (fi ) = V (hfi i). Therefore the ideal I = hfi i contains 1, and this
meansTthat there is some finite subset f1 , . . . , fnSof I such that hf1 , . . . , fn i = R.
n n
Thus i=1 V (fi ) = ∅, or equivalently, Spec R = i=1 U (fi ). 

3. Properties of the spectrum: connectedness


Lemma 13.4. Let E(R) be the set of idempotents in a ring R. Then:
a) If e, f ∈ E(R), then
(i) e∗ := 1 − e ∈ E(R);
(ii) e ∧ f = ef ∈ E(R);
(iii) e ∨ f = (e∗ ∧ f ∗ )∗ = e + f − ef ∈ E(R).
b) (E, ∧, ∨, ∗) is a Boolean algebra.
Exercise 13.5. Prove it.
Exercise 13.6. For a topological space X, let Clopen(X) be the family of clopen
subsets of X.
a) Show: Clopen(X) is a Boolean subalgebra of 2X .
b) Show: if X is compact, then Clopen(X) is the characteristic algebra of X.
Theorem 13.5. Let R be a ring.
a) If e ∈ E(R) is an idempotent, then U (e) is a clopen subset of Spec R.
b) The map E(R) → Clopen(Spec R) given by
e 7→ U (e)
is an isomorphism of Boolean algebras.
Proof. a) If e is idempotent then e(1 − e) = 0, so every p ∈ Spec R contains
exactly one of e and 1 − e and it follows that
U (e) = V (1 − e), U (1 − e) = V (e).
Thus U (e) is clopen.
Jacobson, p. 406. 
Corollary 13.6. For a nonzero ring R, the following are equivalent:
(i) R is connected: E(R) = {0, 1}.
(ii) Spec R is connected.
Exercise 13.7. Prove it.
For the zero ring, we take the convention that it is not connected. This is compatible
with the convention that the empty topological space is not connected.

4. Properties of the spectrum: separation and specialization


For the reader’s convenience we briefly recall the “lower” separation axioms:

A topological space X is Kolmogorov – or T0 – if for any distinct points x, y ∈ X,


the system of neighborhoods Nx and Ny do not coincide. In plainer language,
either there exists an open set U containing x and not containing y, or conversely.

A topological space X is separated – or T1 – if for any distinct points x, y ∈ X,


there exists both an open set U containing x and not y and an open set V contain-
ing y and not x. A space is separated iff all singleton sets {x} are closed iff for all
240 13. Spec R AS A TOPOLOGICAL SPACE

T
x ∈ X, U ∈Nx U = {x}.

A topological space X is Hausdorff – or T2 – if for any distinct points x, y ∈ X,


there exist open neighborhoods U of x and V of y with U ∩ V = ∅. A space is
Hausdorff iff for all x ∈ X, the intersection of all closed neighborhoods of x is {x}.

Easily Hausdorff implies separated implies Kolmogorov. In a general topology


course one learns that neither of the converse implications holds in general. On
the other hand most of the spaces one encounters in analysis and geometry are
Hausdorff, and certainly are if they are Kolmogorov. We are about to see that yet
a third state of affairs transpires when we restrict attention to spectra of rings.

Let X be a topological space. We define a relation 7→ on X by decreeing that


for x, y ∈ X, x 7→ y iff y lies in the closure of the singleton set {x}. This relation
is called specialization, and we read x 7→ y as “x specializes to y”.

The reader who is familiar with topology but has not seen the specialization relation
before will find an explanation in part f) of the following exercise.
Exercise 13.8. a) Show: x 7→ y iff Nx ⊂ Ny .
b) Show that specialization satisfies the following properties:
(i) reflexivity: x 7→ x; (ii) transitivity x 7→ y, y 7→ z =⇒ x 7→ z.
A relation R with these properties is called a quasi-ordering. A partial ordering
is a quasi-ordering with the additional axiom of anti-symmetry: xRy, yRx =⇒
x = y.
c) Show that specialization is a partial ordering on X iff X is Kolmogorov.
d) Show that a point y is closed2 iff y 7→ x =⇒ x = y.
e) A point x for which x 7→ y holds for all y ∈ X is called generic. Give an
example of a topological space in which every point is generic.
f ) Show that X is separated iff x 7→ y =⇒ x = y.
Exercise 13.9. Let X be a set endowed with a quasi-ordering R. Define a new
relation x ≡ y if x R y and y R x.
a) Show: ≡ is an equivalence relation on X.
b) Write X 0 for the set of ≡ equivalence classes, and let q : X → X 0 be the natural
map – i.e., x 7→ {y ∈ X | y ≡ x}. Show that the relation R descends to a relation
≤ on X 0 : i.e., for s1 , s2 ∈ X 0 , then by choosing x1 ∈ s1 , x2 ∈ s2 and putting
s1 ≤ s2 ⇐⇒ x1 R x2 ,
the relation ≤ is well-defined independent of the choices of x1 and x2 . Show that
moreover ≤ is a partial ordering on X 0 .
c) Let X be a topological space and R be the specialization relation. Endowing
X 0 with the quotient topology via q, show that the induced relation ≤ on X 0 is
the specialization relation on X 0 , and accordingly by the previous exercise X 0 is a
Kolmogorov space. If it pleases you, show that q : X → X 0 is universal for maps
from X into a Kolmogorov space Y , hence X 0 (or rather, q : X → X 0 ) can be
regarded as the Kolmogorov quotient of X.

2Strictly speaking we mean {y} is closed, but this terminology is common and convenient.
4. PROPERTIES OF THE SPECTRUM: SEPARATION AND SPECIALIZATION 241

1
Exercise 13.10. Let (X, µ) be a measure R space, and let L be the 1space of
all measurable functions f : X → R with X |f |dµ < ∞. For f ∈ L , define
||f || := X |f |dµ, and for  > 0, put B(f, ) = {g ∈ L1 | ||g − f || < }. Show that
R

the B(f, )’s form a base for a topology on L1 , but that this topology is, in general,
not Kolmogorov. Show that the Kolmogorov quotient is precisely the usual Lebesgue
space L1 , whose elements are not functions but classes of functions modulo µ a.e.
equivalence.
Proposition 13.7. For any ring R, the spectrum Spec R is a Kolmogorov
space. Indeed, for p, q ∈ Spec R we have p 7→ q iff p ⊂ q.
Proof. For prime ideals p and q we have
p 7→ q ⇐⇒ q ∈ {p} = {f ∈ Spec R | f ⊃ p} ⇐⇒ p ⊂ q.
Thus the specialization relation is just containment of ideals, which certainly sat-
isfies antisymmetry: p ⊂ q, p ⊂ q =⇒ p = q. Now apply Exercise 13.9c). 
Theorem 13.8. For a commutative ring R, the following are equivalent:
(i) R/ nil R is absolutely flat, i.e., every R/ nil R-module is flat.
(ii) R has Krull dimension zero.
(iii) Spec R is a separated space.
(iv) Spec R is a Hausdorff space.
(v) Spec R is a Boolean space.
Proof. (i) ⇐⇒ (ii) This is Theorem 7.26.
(ii) ⇐⇒ (iii): A space is separated iff all of its singleton sets are closed. But if p
is prime, V (p) consists of all primes containing p, so V (p) = {p} iff p is maximal.
Certainly (v) =⇒ (iv) =⇒ (iii).
(i) =⇒ (v): Since Spec R = Spec(R/ nil R), we may well assume that R itself is
absolutely flat. Let p and q be distinct prime ideals; since both are maximal, there
exists an element f ∈ p \ q. By Proposition 3.100, there is an idempotent e with
(e) = (f ), and therefore e ∈ p \ q. Then D(1 − e), D(e) is a separation of Spec R.
More precisely, D(e) ∩ D(1 − e) = D(e(1 − e)) = D(e − e2 ) = D(0) = ∅, whereas
for any prime ideal p, since 0 = e(1 − e) ∈ p, we must have e ∈ p or 1 − e ∈ p. By
construction, p ∈ D(1 − e), q ∈ D(e). This shows Spec`R is Hausdorff, and more:
given points P 6= Q of X, we found a separation X = U V with P ∈ U, Q ∈ V , so
X is zero-dimensional. By Proposition 13.3, every ring has quasi-compact spectrum,
so Spec R is Hausdorff, zero-dimensional and quasi-compact, i.e., Boolean. 
Exercise 13.11. a) Let R be a product of fields. Show: Spec R is a Boolean
space.
Q{Ri }i∈I be a family of rings, each of which has Krull dimension 0, and put
b) Let
R = i Ri . Must Spec R be Boolean?
4.1. Gelfand Rings.
Theorem 13.9. a) For a ring R, the following are equivalent:
(i) Every prime ideal of R is contained in a unique maximal ideal.
(ii) For all x ∈ R, there are r, s ∈ R such that (1 + rx)(1 + s(1 − x)) = 0.
(iii) Spec R is quasi-normal: two disjoint closed sets can be separated by two disjoint
open sets.
(iv) MaxSpec R is a retract of Spec R. A ring satisfying these equivalent conditions
is called a Gelfand ring.
242 13. Spec R AS A TOPOLOGICAL SPACE

Proof. COMPLETE ME! 


Exercise 13.12. Show: the class of Gelfand rings is closed under quotients,
localizations and direct products.
Exercise 13.13. (E. Wofsey3) Show: if R is a Gelfand ring, then MaxSpec R
is a deformation retract of Spec R. In particular, the two spaces are homotopy
equivalent.
Theorem 13.10. For a ring R, the following are equivalent:
(i) R/J(R) is Gelfand.
(ii) MaxSpec R is compact.
(iii) MaxSpec R is Hausdorff.
Proof. COMPLETE ME! 
A clean ring (also called an exchange ring) is a ring in R in which for all x ∈ R
there is an idempotent e ∈ R such that x + e ∈ R× .
Exercise 13.14. a) Show: every Boolean ring is a clean ring.
b) Show: every clean ring is a Gelfand
   (Hint: for x ∈ R, let e be an idempotent
ring.
such that e − x ∈ R× . Show: 1+ x
e−x 1− 1−x
e−x = 0.)
c) Show: every local ring is a clean ring. If Spec R is connected, then R is clean iff
it is local. d) Show: a Gelfand domain must be local.
Exercise 13.15. a) Show: for a topological space X, C(X) is a Gelfand ring.
(Hint: for f ∈ C(X), max(f, 1 − f ) ∈ C(X)× .).
b) Prove Corollary 5.15.
c) Show: C(X) need not be clean, but if X is Boolean (i.e., compact and totally
disconnected) then C(X) is clean.

5. Irreducible spaces
A topological space is irreducible if it is nonempty and if it cannot be expressed
as the union of two proper closed subsets.
Exercise 13.16. Show: for a Hausdorff topological space X, the following are
equivalent:
(i) X is irreducible.
(ii) #X = 1.
Proposition 13.11. For a topological space X, the following are equivalent:
(i) X is irreducible.
(ii) Every finite intersection of nonempty open subsets (including the empty inter-
section!) is nonempty.
(iii) Every nonempty open subset of X is dense.
(iv) Every open subset of X is connected.
Exercise 13.17. Prove Proposition 13.11.
Proposition 13.12. Let X be a nonempty topological space.
a) If X is irreducible, every nonempty open subset of X is irreducible.
b) If a subset Y of X is irreducible, so is its closure Y .
3Cf. http://math.stackexchange.com/questions/1586745
5. IRREDUCIBLE SPACES 243

c) If {Ui } is an open covering of X such that Ui ∩ Uj 6= ∅ for all i, j and each Ui


is irreducible, then X is irreducible.
d) If f : X → Y is continuous and X is irreducible, then f (X) is irreducible in Y .
Proof. a) Let U be a nonempty open subset of X. By Proposition 13.11, it
suffices to show that any nonempty open subset V of U is dense. But V is also a
nonempty open subset of the irreducible space X.
b) Suppose Y = A ∪ B where A and B are each proper closed subsets of Y ; since
Y is itself closed, A and B are closed in X, and then Y = (Y ∩ A) ∪ (Y ∩ B). If
Y ∩ A = Y then Y ⊂ A and hence Y ⊂ A = A, contradiction. So A is proper in Y
and similarly so is B, thus Y is not irreducible.
c) Let V be a nonempty open subset of X. Since the Ui ’s are a covering of X, there
exists at least one i such that V ∩ Ui 6= ∅, and thus by irreducibility V ∩ Ui is a
dense open subset of Ui . Therefore, for any index j, V ∩ Ui intersects the nonempty
open subset Uj ∩ Ui , so in particular V intersects every element Uj of the covering.
Thus for all sets Ui in an open covering, V ∩ Ui is dense in Ui , so V is dense in X.
d) If f (X) is not irreducible, there exist closed subsets A and B of Y such that
A ∩ f (X) and B ∩ f (X) are both proper subsets of f (X) and f (X) ⊂ A ∪ B. Then
f −1 (A) and f −1 (B) are proper closed subsets of X whose union is all of X. 

Proposition 13.13. Let R be a ring. Let V ⊂ Spec R be a Zariski-closed


subset, so V = V (I) for a unique radical ideal I. The following are equivalent:
(i) The subspace V (I) is irreducible.
(ii) The ideal I is prime.
Proof. ¬ (ii) =⇒ ¬ (i): Let a, b ∈ R such that ab ∈ R and a, b ∈ R \ I.
Then I contains neither rad(a) nor rad(b), so V (I) is contained in neither V (a) nor
V (b), but V (a) ∪ V (b) = V (ab) ⊃ V (I). Thus V (a) ∩ V (I) and V (b) ∩ V (I) are two
proper closed subsets whose union is V (I), so V (I) is reducible.
(ii) =⇒ (i): Suppose p is a prime ideal. We claim that V (p) is irreducible. If
not, there are ideals I and J such that V (I) and V (J) are both proper subsets of
V (p) and V (p) = V (I) ∪ V (J) = V (IJ). But then p = rad(IJ) ⊃ IJ and since p
is prime this implies p ⊃ I or p ⊃ J. WLOG, suppose p ⊃ I; then V (p) ⊂ V (I), so
that V (I) is not proper in V (p), contradiction. 

Let x be a point of a topological space, and consider the set of all irreducible sub-
spaces of X containing x. (Since {x} itself is irreducible, this set is nonempty.) The
union of a chain of irreducible subspaces being irreducible, Zorn’s Lemma says that
there exists at least one maximal irreducible subset containing x. A maximal irre-
ducible subset (which, by the above, is necessarily closed) is called an irreducible
component of X. Since irreducible subsets are connected, each irreducible com-
ponent lies in a unique connected component. It is not true in general that the
connected component of x is the union of the irreducible components containing x:
the way that irreducible components are combined to form connected components
is slightly more complicated and explored in Exercise 13.22.

However, unlike connected components, it is possible for a given point to lie in


more than one irreducible component. We will see examples shortly.

In the case of the Zariski topology on Spec R, it follows from Proposition 13.13
244 13. Spec R AS A TOPOLOGICAL SPACE

that the irreducible components of Spec R are the subsets V (p) for a minimal
prime ideal p. Above we showed that every nonempty topological space has at
least one irreducible component, so this argument shows that every nonzero ring
admits minimal primes. Thus we have deduced a commutative algebraic result as a
consequence of a topological result. However we showed this earlier in Proposition
4.33 and the two proofs are essentially the same Zorn’s Lemma argument. A more
interesting topological argument is coming up soon.

6. Noetherianity
6.1. Noetherian topological spaces. We now introduce a property of topo-
logical spaces which, from the standpoint of conventional geometry, looks com-
pletely bizarre:
Proposition 13.14. For a topological space X, the following are equivalent:
(i) Every ascending chain of open subsets is eventually constant.
(ibis) Every descending chain of closed subsets is eventually constant.
(ii) Every nonempty family of open subsets has a maximal element.
(iibis) Every nonempty family of closed subsets has a minimal element.
(iii) Every open subset is quasi-compact.
(iv) Every subset is quasi-compact.
A space satisfying any (and hence all) of these conditions is called Noetherian.
Proof. The equivalence of (i) and (ibis), and of (ii) and (iibis) is immediate
from taking complements. The equivalence of (i) and (ii) is a general property of
partially ordered sets discussed in section X.X above.
(i) ⇐⇒ (iii): Assume (i), let U be any open set in X and let {Vj } be an open
covering of U . We assume for a contradiction that there is no finite subcovering.
Choose any j1 and put U1 := Vj1 . Since U1 6= U , there exists j2 such that U1 does
not contain Vj2 , and put U2 = U1 ∪ Vj2 . Again our assumpion implies that U2 ) U ,
and continuing in this fashion we will construct an infinite properly ascending chain
of open subsets of X, contradiction. Conversely, assumeS(iii) and let {Ui }∞ i=1 be an
infinite properly ascending chain of subsets. Then U = i Ui is not quasi-compact.
Obviously (iv) =⇒ (iii), so finally we will show that (iii) =⇒ (iv). Suppose
that Y ⊂ X is not quasi-compact, and let {Vi }i∈I be a covering of Y by relatively
open subsets without
S a finite subcover. We may write each Vi as Ui ∩Y with Ui open
in Y . Put U = i Ui .SThen, since U is quasi-compact,Sthere exists a S finite subset
J ⊂ I such that U = j∈J Uj , and then Y = U ∩ Y = j∈J Uj ∩ Y = j∈J Vj . 
Corollary 13.15. A Noetherian Hausdorff space is finite.
Proof. In a Hausdorff space every quasi-compact subset is closed. Therefore,
using the equivalence (i) ⇐⇒ (iv) in Proposition 13.14, in a Noetherian Hausdorff
space every subset is closed, so such a space is discrete. But it is also quasi-compact,
so it is finite. 
Proposition 13.16. Let X be a topological space.
a) The following are equivalent:
(i) There is a finite partition of X into connected clopen subsets.
(ii) The space X has finitely many connected components.
(iii) The space X has finitely many clopen subsets.
b) The equivalent conditions of part a) hold when X is Noetherian.
6. NOETHERIANITY 245

Proof. a) (i) =⇒ (ii): Whenever a topological space admits a partition into


connected clopen subsets, these subsets are the connected components of X.
(ii) =⇒ (i): In any topological space, each connected component is closed. If
there are only finitely many connected components then the complement of each
component is a finite union of closed sets, so each connected component is clopen.
(i) =⇒ (iii): Intersecting any clopen set with a connected component C(x) gives
either the empty set or C(x), and the result follows easily from this.
(iii) =⇒ (i): A partition of X into clopen sets is maximal iff each clopen set in
the partition is connected, so in any nonmaximal partition we can partition one of
the clopen sets in the partition into two clopen subsets, increasing the number of
clopen sets in the partition by 1. So if X has finitely many clopen subsets then
this process, starting with {X}, must terminate after finitely many steps, yielding
a finite partition into connected clopen subsets.
b) Suppose X is Noetherian, and let F be the family of closed subsets of X which
have infinitely many connected components. If F is nonempty, then by since X is
Noetherian ` F has a minimal element Y . Then Y is nonempty and disconnected,
so Y = Y1 Y2 with Y1 , Y2 nonempty. By minimality of Y , the proper closed
subsets Y1 and Y2 each have finitely many connected components, hence so does Y :
contradiction. Applying this to X itself, we get that X has finitely many connected
components. 
Just to be sure: if a topological space X has infinitely many connected components,
it may or may not admit a partition into clopen subsets. This does hold if X is
a paracompact topological manifold. Notice though that this holds for a totally
disconnected space iff the space is discrete, and there are many totally disconnected
spaces that are not discrete, e.g. any infinite boolean space or the rational numbers
as a subspace of the real numbers.
Proposition 13.17. Let X be a Noetherian topological space. Sn
a) There are finitely many closed irreducible subsets {Ai }ni=1 such that X = i=1 Ai .
b) Starting with any finite family {Ai }ni=1 as in part a) and eliminating all redun-
dant sets – i.e., all Ai such that Ai ⊂ Aj for some j 6= i – we arrive at the
set of irreducible components of X. In particular, the irreducible components of a
Noetherian space are finite in number.
Proof. a) Let X be a Noetherian topological space. We first claim that X
can be expressed as a finite union of irreducible closed subsets. Indeed, consider
the collection of closed subsets of X which cannot be expressed as a finite union
of irreducible closed subsets. If this collection is nonempty, then by Proposition
13.14 there exists a minimal element Y . Certainly Y is not itself irreducible, so is
the union of two strictly smaller closed subsets Z1 and Z2 . But Z1 and Z2 , being
strictly smaller than Y , must therefore be expressible as finite unions of irreducible
closed subsets and therefore so also can Y be so expressed, contradiction.
b) So write
X = A1 ∪ . . . ∪ An
where each Ai is closed and irreducible. If for some i 6= j we have Ai ⊂ Aj , then
we call Ai redundant and remove it from our list. After a finite number of such
removals, we may assume that the above finite covering of X by closed irreducibles is
irredundant in the sense that there are no containment relationsSn between distinct
Ai ’s. Now let Z be any irreducible closed subset. Since Z = i=1 (Z ∩ Ai ) and Z is
246 13. Spec R AS A TOPOLOGICAL SPACE

irreducible, we must have Z = Z ∩ Ai for some i, i.e., Z ⊂ Ai . It follows that the


“irredundant” Ai ’s are precisely the maximal irreducible closed subsets, i.e., the
irreducible components. 
6.2. Applications to Noetherian rings.
Proposition 13.18. For a ring R, the following are equivalent:
(i) R satisfies the ascending chain condition on radical ideals.
(ii) Spec R is a Noetherian space.
In particular if R – or even Rred = R/ nil(R) – is a Noetherian ring, then Spec R
is a Noetherian space.
Proof. Since I 7→ V (I) gives a bijection between radical ideals and Zariski
closed subsets, (ACC) on radical ideals is equivalent to (DCC) on closed subsets.
Evidently these conditions occur if R is itself Noetherian, or, since Spec R is canon-
ically homeomorphic to Spec Rred , if Rred is Noetherian. 
Corollary 13.19. Let I be a proper ideal in a Noetherian ring R. The set
of prime ideals p which are minimal over I (i.e., minimal among all prime ideals
containing I) is finite and nonempty.
Exercise 13.18. Prove Corollary 13.19.
Exercise 13.19. QLet I be an infinite set, and for all i ∈ I, let ri be a nonzero
ring. Show: the ring i∈I ri is not Noetherian.
The following result shows that every Noetherian ring is a finite product of con-
nected Noetherian rings in an essentially unique way.
Theorem 13.20. Let R be a nonzero Noetherian ring.
a) There is a unique n ∈ Z+ for which there are I1 , . . . , In ∈ I(R) such that:
(i) For all 1 ≤ i ≤ n the ring R/Ii is connected.
(ii) For all 1 ≤
Tni ≤ n there is an idempotent ei ∈ R such that Ii = hei i.
(iii) We have i=1 Ii = (0) and Ii + Ij = Rf or all 1 ≤ i 6= j ≤ n. Thus
we have a canonical isomorphism
n

Y
(31) R→ R/Ii .
i=1

b) Let N ∈ Z+ and let r1 , . . . , rN be nonzero rings such that R ∼ =


QN
i=1 ri . Then N ≤ n. Moreover, if each ri is connected, then we
have N = n and after reordering the ri ’s we have for all 1 ≤ i ≤ n
an R-algebra isomorphsim ri ∼= R/Ii .
`n
Proof. a) By Proposition 13.16 there is n ∈ Z+ such that Spec R = i=1 Xi
with each Xi nonempty connected, and the Boolean algebra Clopen Spec R has
order 2n . For 1 ≤ i ≤ n, put Yi := Spec R \ Xi = 1≤j≤n, j6=i Xj ∈ Clopen Spec R.
`
By Theorem 13.5, for all 1 ≤ i ≤ n there is an idempotent ei ∈ R such that
Yi = U (ei ). We know that e 7→ U (e) gives an isomorphism of Boolean algebras, so
for all 1 ≤ i 6= j ≤ n we have
n
\ n
\
U (e1 · · · en ) = U (e1 ∧ . . . ∧ en ) = U (ei ) = Yi = ∅ = U (0),
i=1 i=1
6. NOETHERIANITY 247

Tn
so e1 · · · en = 0. If x ∈ i=1 Ii , then for all 1 ≤ i ≤ n we have x = ei xi for some
2
Tin∈ R and thus ei x = ei xi = ei xi = x. It follows that x = e1 · · · en x = 0, so
x
i=1 Ii = (0). For 1 ≤ i 6= j ≤ n we have
U (ei ∧ ej ) = U (ei ) ∪ U (ej ) = Yi ∪ Yj = X = U (1),
so ei ∧ ej = 1. Thus
1 = ei ∧ ej = ei + ej − ei ej ∈ Ii + Ij .
So we may apply the Chinese Remainder Theorem to get (31). Moreover we have
Spec R/Ii = Spec R/hei i = V (ei ) = Spec R \ U (ei ) = Spec R \ Yi = Xi .
Finally, (31) implies that n must be the number of connected components of Spec R.
b) Suppose there are nonzero rings r1 , . . . , rN such that R ∼
QN
= i=1 ri . Each ri is a
quotient of R hence Noetherian, so by part a) is itself a finite product of connected
nonzero rings. So if N > n then this would express R as a product of more than
n connected nonzero rings, contradicting part a). Let us now assume that each ri
is connected. Again, this forces N = n. For 1 ≤ i ≤ n let Ji be the kernel of the
natural map R → ri , so we get an R-algebra isomorphism ri ∼ = R/Ji . By Exercise
4.13 we have Ji + Jj = R for all 1 ≤ i 6= j ≤ n. Thus for all 1 ≤ i 6= j ≤ n we have
n
[ n
[ n
\
Spec R/Ji = V (Ji ) = V ( Ji ) = V (0) = Spec R
i=1 i=1 i=1
and
Spec R/Ji ∩ Spec R/Jj = V (Ji ) ∩ V (Jj ) = V (Ji + Jj ) = V (R) = ∅.
Thus the V (Ji )’s are precisely the connected components of Spec R, so after re-
ordering we have V (Ji ) = V (Ii ) and thus rad Ji = rad Ii for all 1 ≤ i ≤ n. Both Ii
and Ji are ideals genearted by an idempotent element: this was the construction
of Ii , while Ji = {(r1 , . . . , ri−1 , 0, ri+1 , . . . , rn ) | rj ∈ rjQ
} so it is generated by the
n
element of R corresponding to (1, 1, . . . , 0, 1, . . . , 1) ∈ i=1 ri . Since both Ii and
Ji are finitely generated, by Proposition 4.15g) this implies that there is M ∈ Z+
such that IiM ⊂ Ji and JiM ⊂ Ii . It follows that Ji = Ii for all 1 ≤ i ≤ n. 
The above result continues to hold under weaker hypotheses.
Exercise 13.20. Let R be a nonzero ring, and let κ be the number of connected
components of Spec R.
a) Show that the following are equivalent:
(i) We have κ < ℵ0 .
(ii) The ring R has 2κ < ℵ0 idempotents.
(iii) The ring R is a finite product of connected rings.
b) When the equivalent conditions of part a) hold, if R is a product of α
nonzero rings, then α ≤ κ. If R is a product of β connected rings, then
β = κ. Moreover if we have connected rings r1 , . . . , rκ , s1 , . . . , sκ such that
κ κ
ri ∼
=R∼
Y Y
= si ,
i=1 i=1
then after permuting the indices we have R-algebra isomorphisms ri ∼
= si
for all 1 ≤ i ≤ κ.
Exercise 13.21. Let R1 , R2 , R3 be nonzero rings such that R1 × R2 ∼
= R1 × R3 .
248 13. Spec R AS A TOPOLOGICAL SPACE

a) Show: if each Ri is Noetherian (or indeed is a finite product of connected


rings), show that R2 ∼
= R3 .
b) Give an example to show that we need not have R2 ∼ = R3 in general.
Exercise 13.22. Let X be a Noetherian topological space, and let x ∈ X.
a) Show: every irreducible component containing x is contained in the connected
component of x. Deduce: every connected component is a finite union of irreducible
components.
b) Show: the union of the irreducible components containing x may be a proper
subset of the connected component of x.
c) For two irreducible components of X, write Ai ∼ Aj if Ai ∩ Aj 6= ∅. Show: ∼
need not be transitive. Let ≈ be the transitive closure of ∼: explicitly, A ≈ B if
there is a finite chain A ∼ A1 ∼ . . . ∼ An ∼ B. Show: ≈ is an equivalence relation
on the set of irreducible components of X and the unions of the components in any
one equivalence class are precisely the connected components of X.

7. Krull Dimension of Topological Spaces


The Krull dimension dim X of a topological space is the supremum of lengths of
chains of irreducible closed subsets of X. This is a cardinal number. In a ring R we
have a bijective correspondence between prime ideals of R and irreducible closed
subspaces of Spec R, which gives
dim R = dim Spec R.
Example 13.21. Let X be the topological space with underlying set Z+ and
with nonempty open sets
U (n) = {m ∈ Z+ | m ≥ n}
as n ranges over all positive integers. Then families of nonempty open sets are
subsets A ⊂ Z+ . If A 6= ∅ then n∈A U (n) = U (min
S
indexed by T T A). If A is
finite, then n∈A U (n) = U (max A), whereas if A is infinite then n∈A U (n) = ∅.
Thus X is an Alexandroff topological space: the family of open sets is closed under
arbitrary unions and intersections.
Since every nonempty open subset of X is cofinite, X is Noetherian. For n ∈
Z+ , the proper closed sets containing x are the intervals [1, m] = {1, . . . , m} with
n ≤ m, so n = [1, n]. Thus every nonempty closed subset of X is irreducible, and
` subset has a generic point iff it is proper, and dim X = ℵ0 .
an irreducible closed
Let X̃ = X {η}, where η is some point not in X. We define a nonempty
subset of X̃ to be open if it is of the form U (n) ∪ {η} for some n ∈ Z+ . This gives a
topology on X̃. Since every nonempty open subset of X̃ contains η, we have η = X̃.
Let ι : X ,→ X̃. The map ι−1 gives an order-preserving bijection from the open
subsets of X̃ to the open substs of X. Thus X̃ is Noetherian and dim X̃ = ℵ0 .

8. Jacobson spaces
Let X be a topological space. We denote by X0 the subset of closed points of X.
We endow X0 with the subspace topology.
Exercise 13.23. Let X be a topological space.
a) Show: if X is finite, then X0 is closed in X.
b) Exhibit a topological space X for which X0 is not closed in X.
(Suggestion: take X = Spec R for a suitable ring!)
8. JACOBSON SPACES 249

A subset Y ⊂ X is locally closed if it is the intersection of an open set with a


closed subset. A subset Z ⊂ X is strongly dense if Z ∩ Y 6= ∅ for all nonempty
locally closed subsets Y ⊂ X. A topological space X is Jacobson if X0 is strongly
dense in X.

Notice that X is separated iff X0 = X; such spaces are certainly Jacobson. The
concept is only interesting when not all points are closed.
Lemma 13.22. For a topological space X, the following are equivalent:
(i) X0 is strongly dense in X.
(ii) For all closed subsets Z ⊂ X, we have Z = Z ∩ X0 .
(iii) For all x ∈ X, we have {x} = {x} ∩ X0 .
A space satisfying these equivalent properties is called a Jacobson space.
Exercise 13.24. Prove it.
Exercise 13.25. a) Let X be at topological space, and let {Ui }i∈I be an open
cover of X. Show: X is Jacobson iff Ui is Jacobson for all i.
b) Suppose X is Jacobson and Y ⊂ X is a union of locally closed subspaces of X.
Show: Y is Jacobson.
c) Show: a finite Jacobson space is separated (hence discrete).
Exercise 13.26. For a topological space X, we define the Jacobson subspace

J(X) = {x ∈ X | {x} = {x} ∩ X0 }.


a) Show: X is Jacobson iff J(X) = X.
b) Show: J(X) is a Jacobson space.
Lemma 13.23. Let X be a Jacobson topological space, and let ι : X0 → X be
the inclusion map.
a) The map Y ⊂ Xmapstoι−1 (Y ) = Y ∩ X0 s a bijection from the closed subspaces
of X to the closed subspaces of X0 .
b) A closed subset Y ⊂ X is irreducible iff Y ∩ X0 is irreducible.
c) We have dim X0 = dim X.
Proof. a) Certainly if Y ⊂ X is closed in X, then Y ∩ X0 is closed in X0 , and
by definition of the subspace topology every closed subset of X0 is of the form Y ∩X0
for some closed Y ⊂ X. Suppose Y1 , Y2 are closed in X and Y1 ∩ X0 = Y2 ∩ X0 . If
Y1 6= Y2 , then (Y1 \ Y2 ) ∪ (Y2 \ Y1 ) is a nonempty locally closed set, so it meets X0 :
contradiction.
b), c) Left to the reader. 

Proposition 13.24. For a ring R, the following are equivalent:


(i) R is a Jacobson ring.
(ii) Spec R is a Jacobson space.
Proof. Let p ∈ Spec R. Then p = {q ∈ Spec R | q ⊃ p} and
\
{p} ∩ MaxSpec R = m,
m∈MaxSpec R, m⊃p

so the equivalence follows from Lemma X.X. 


250 13. Spec R AS A TOPOLOGICAL SPACE

For a ring R, we define the Jacobson spectrum JSpec R to be the set of all prime
ideals which are intersections of maximal ideals.4 Thus MaxSpec R ⊂ JSpec R ⊂
Spec R, and JSpec R = Spec R iff R is Jacobson. We endow JSpec R with the
topology it receives as a subspace of Spec R. Since JSpec R consists precisely of the
prime ideals of R which lie in the closure of the set of maximal ideals containing
them, we have that
JSpec R = J(Spec R),
i.e., JSpec R is the Jacobson subspace of Spec R. In particular, JSpec R is a Jacob-
son space.

A topological space X is sober if for every irreducible closed subspace Y of X,


there exists a unique point y ∈ Y such that Y = {y}. Equivalently, a sober space
is one for which every irreducible closed subset has a unique generic point.

Exercise 13.27. a) Show that any Hausdorff space is sober.


b) Show that a sobser space is Kolmogorov.
c) Show that the cofinite topology on an infinite set is separated but not sober.

A map f : (X, τX ) → (Y, τY ) of topological spaces is a quasi-homeomorphism if


it is continuous and f −1 : V ∈ τY 7→ f −1 (V ) ∈ τX is a bijection.

Exercise 13.28. (Sobrification) For a topological space X, let X + denote the


set of irreducible closed subspaces of X. If Y ⊂ X is closed, then Y + ⊂ X + .
Consider the family C = {Y + | Y ⊂ X is closed } of subsets of X + .
a) Show: C contains ∅ and X + and is closed under finite unions and arbitrary
intersections, thus forms the closed sets for a unique topology on X + .
Define a map j : X → X + by j : x 7→ {x}.
b) Show: if Y ⊂ X is closed, then j −1 (Y + ) = Y . Deduce: V 7→ V + is a bijection
from the family of closed subsets of X to the family of closed subsets of X + , with
inverse bijection V + 7→ f −1 (V + ) = V . In particular, j is quasi-homeomorphism.
c) Show: X has the initial topology coming from the map j : X → X + , i.e., the
coarsest topology that makes j continuous.
d) Show: if V ⊂ X is closed, then V is irreducible iff V + is irreducible, so V 7→ V +
gives a bijection from the closed irreducible subsets of X to the closed irreducible
subsets of X + . Deduce: X + is sober.
e) Show: the map j : X → X + is the universal continuous function from X into a
sober topological space: if f : X → Y is continuous and Y is sober, then there is a
unique F : X + → Y such that f = F ◦ j. We say that j : X → X + ( and, by a
standard abuse of terminology, X + itself ) is the sobrification of X.
f ) Show: j is injective iff X is Kolmogorov and j is surjective iff every irreducible
closed subset of X has a generic point (quasi-sober).
g) Show: the following are equivalent:
(i) X is sober.
(ii) j is a homeomorphism.
(iii) j is bijection.
(iv) j is a closed injection.

4The Jacobson spectrum was introduced by R.G. Swan [Sw67].


8. JACOBSON SPACES 251

Proposition 13.25. Let f : X → Y be a quasi-homeomorphism.


a) If X is Kolmogorov, then f is injective.
b) If X is sober and Y is Kolmogorov, then f is a homeomorphism.
Proof. a) Seeking a contradiction, we suppose there are x1 6= x2 in X such
that f (x1 ) = f (x2 ). After interchanging x1 and x2 if necessary, we may assume
there is an open subset U ⊂ X containing x1 and not x2 . Let V ⊂ Y be open such
that q −1 (V ) = U . Then q(x1 ) ∈ V and q(x2 ) ∈
/ V : contradiction.
b) It suffices to show f is bijective, since a bijective quasi-homeomorphism is a
homeomorphism. By part a), f is injective. Let y ∈ Y . Then f −1 ({y}) is irreducible
and closed in the sober space X, so it has a generic point x. Thus
{x} ⊂ f −1 ({f (x)}) ⊂ f −1 ({y}) = {x},
so f −1 ({f (x)}) = f −1 ({y}). Since f is quasi-homeomorphism, we have {f (x)} =
{y}, and since Y is Kolmogorov we conclude f (x) = y. 
Corollary 13.26. If f : X → Y is a quasi-homeomorphism and Y is sober,
then f is the sobrification of X.
Proof. The universal property of the sobrification j : X → X + gives us a
factorization f = F ◦j for a continuous map F : X + → Y . We have f −1 = j −1 ◦F −1 .
Since f −1 and j −1 are each bijections between lattices of open sets, so is F −1 . Thus
F is a quasi-homeomorphism of sober spaces, hence a homeomorphism. 
Example 13.27. Let X be a topological space which is Kolmogorov but not
sober: e.g. take X to be an inifnite endowed with the cofinite topology. Let j :
X → X + be the sobrification, so j is a quasi-homeomorphism. There is no quasi-
homeomorphism f : X + → X: indeed, the previous result implies that f would have
to be a homeomorphism and thus X would be sober, contradiction.
So beware: as a relation, quasi-homeomorphism is not symmetric, hence not an
equivalence relation. However, if ≡ is the equivalence relation on topological spaces
generated by quasi-homeomorphism, then it is easy to see that X ≡ Y iff X + and
Y + are homeomorphic.
We come now to the following result, which is the main payoff of the material of
this section and will be used in our treatment of the Forster-Swan Theorem.
Theorem 13.28. For any ring R, the inclusion map ι : MaxSpec R ,→ JSpec R
is the sobrification of MaxSpec R.
Proof. We leave it to the reader to check that the argument that shows
Spec R is sober carries over to show the sobriety of JSpec R. By the above corol-
lay it is enough to show that ι is a quasi-homeomorphism. Since ι is an inclu-
sion map, certainly ι−1 is surjective on closed sets, so it’s enough to see injectiv-
ity. If Z is a closed subset of JSpec R, then since JSpec R is Jacobson, we have
Z = Z ∩ MaxSpec R = ι−1 (Z), giving the injectivity. 
Corollary 13.29. Let R be a ring.
a) The following conditions are equivalent:
(i) MaxSpec R is Noetherian.
(ii) JSpec R is Noetherian.
(iii) The ascending chain condition holds on intersections of maximal ideals in R.
b) If Spec R is Noetherian, then the conditions of part a) hold.
252 13. Spec R AS A TOPOLOGICAL SPACE

c) The supremum of lengths of chains in JSpec R is the Krull dimension of both


JSpec R and of MaxSpec R.
d) We have dim MaxSpec R ≤ dim Spec R.
Exercise 13.29. Prove it.

9. Hochster’s Theorem
A topological space X is spectral if:
(SS1) X is quasi-compact,
(SS2) X is sober, and
(SS3) The family of quasi-compact open subsets of X is closed under finite inter-
sections and is a base for the topology.

Remark: A Bourbakiste would insist that (SS3) =⇒ (SS1) by taking the empty
intersection. But we will not do so.

Exercise 13.16: Show that a finite space is spectral iff it is T0 .

The following result gives an arguably cleaner characterization of spectral spaces.


Proposition 13.30. . For a topological space X, the following are equivalent:
(i) X is homeomorphic to an inverse limit of finite T0 spaces.
(ii) X is spectral.
Exercise 13.17: Prove Proposition 13.30.
Proposition 13.31. For any ring R, Spec R is spectral.
Exercise 13.30. Prove Proposition 13.31.
(Hint: you will find the needed results in the previous subsections. Especially, use
Proposition 13.13 to prove sobriety.)
For any ring R we endow the set MaxSpec(R) of maximal ideals of R with the
topology it inherits as a subset of Spec(R). When necessary, we describe MaxSpec R
as the “maximal spectrum” of R.
Proposition 13.32. For any ring R, MaxSpec R is separated and quasi-compact.
Exercise 13.31. Prove Proposition 13.32.
Theorem 13.33. (Hochster’s Thesis [Ho69])
a) A spectral topological space is homeomorphic to the prime spectrum of some ring.
b) A separated quasi-compact space is homeomorphic to the maximal spectrum of
some ring.
We do not aspire to give a proof of Theorem 13.33 at this time.
Exercise 13.32. Show: every compact space X is homeomorphic to MaxSpec(C(X)),
where C(X) is the ring of continuous R-valued functions on X.
Exercise 13.33. a) Show: the specialization relation gives an equivalence of
categories between the category of T0 finite spaces and the category of finite partially
ordered sets.
b)* Formulate a generalization of part a) in which T0 finite spaces are replaced by
T0 Alexandroff spaces.
10. RANK FUNCTIONS REVISITED 253

(A topological space is Alexandroff if an arbitrary intersection of closed subsets is


closed.)
Exercise 13.34. Let n ∈ Z+ .
a) Use Hochster’s Thesis and the previous exercise to show: there exists a ring R
with exactly n prime ideals p1 , . . . , pn such that p1 ⊂ p2 ⊂ . . . ⊂ pn .
b) For n = 1, 2, exhibit Noetherian rings with these properties. For n ≥ 3, show:
there is no such Noetherian ring.

10. Rank functions revisited


Theorem 13.34. Let M be a finitely generated module over a ring R.
a) For each n ∈ N, the set
Ur = {p ∈ Spec R | Mp can be generated over Rp by at most r elements}
is open in Spec R.
b) If M is finitely presented (e.g. if R is Noetherian), then the set
UF = {p ∈ Spec R | Mp is a free Rp -module}
is open in Spec R.
Proof. (Matsumura) Suppose Mp = hω1 , . . . , ωr iRp . Each ωi is of the form
mi
si with mi ∈ M and si ∈ R \ p. But since si ∈ Rp× for all i, we also have
hm1 , . . . , mr iRp = Mp . Thus it is no loss of generality to assume that each ωi is
r
the image in Mp of Pan element of M . Let ϕ : R → M be the R-linear map given
by (a1 , . . . , ar ) 7→ i ai ωi , and put C = coker ϕ, whence an exact sequence
Rr → M → C → 0.
Localizing this at a prime q of R gives an exact sequence
Rqr → Mq → Cq → 0.
When q = p we of course have Cq = 0. Moreover, C is a quotient of M hence
a finitely generated R-module, so by Proposition 10.11 its support supp C is a
Zariski-closed set. It follows that there exists an open neighborhood V of p such
that Cq = 0 for all q ∈ V .
b) Suppose that Mp is a free Rp -module with basis ω1 , . . . , ωr . As above it is no
loss of generality to assume that each ωi is the image in Mp of an element of M .
Moreover, as we have also just seen, there exists a basic open neighborhood U (f )
such that for all q ∈ U (f ), the images of ω1 , . . . , ωr in Mq generate Mq as an Rq -
module. Replacing R by Rf and M by Mf we may assume that this occurs for all
q ∈ Spec R. Thus M/hω1 , . . . , ωr iR is everywhere locally zero, so it is locally zero:
M = hω1 , . . . , ωr i. Defining an R-linear map ϕ : Rr → M as above and setting
K = Ker ϕ, we have the exact sequence
0 → K → Rr → M → 0.
Since M is finitely presented, according to Proposition 3.6 K is a finitely generated
R-module. Moreover we have Kp = 0 hence as above Kq = 0 for all q on some open
neighborhood V of p. By construction, for each q ∈ V , the images of ω1 , . . . , ωr in
Mq give an Rq -basis for Mq . 
254 13. Spec R AS A TOPOLOGICAL SPACE

Let M be a finitely generated, locally free module over a ring R. Earlier we defined
the rank function r : Spec R → N. Applying Theorem 13.34a) to the locally free
module M says that the rank function is upper semicontinuous: it can jump up
upon specialization, but not jump down.

We now ask the reader to look back at Theorem 7.29 and see that for a finitely
generated module M over a general ring R, M is projective iff it is locally free and
finitely presented. When R is Noetherian, being finitely presented is equivalent to
being finitely generated, so being projective is the same as being locally free. How-
ever, in the general case we have had little to say about the distinction between
finitely presented and finitely generated modules. Is there some way to rephrase
the subtly stronger property of finite presentation, perhaps a more geometric way?

Indeed there is:


Theorem 13.35. Let M be a finitely generated locally free R-module. the fol-
lowing are equivalent:
(i) The rank function rM : Spec R → N is locally constant.
(ii) M is a projective module.
Proof. (i) =⇒ (ii): By Theorem 7.29, it is enough to show that for all
m ∈ MaxSpec R, there is f ∈ R \ m such that Mf is a free module. Let n = r(m),
and let x1 , . . . , xn be an Rm -basis for Mm . Choose X1 , . . . , Xn ∈ M such that for all
×
i, the image of Xi in Mm is of the form ui xi for uu ∈ Rm . Let u : Rn → M be the
map sending the ith standard basis element ei to Xi . Since M is finitely generated,
by Proposition 7.27 there is f ∈ R \ m such that uf : Rfn → Mf is surjective. It
follows that for all g ∈ R \ m, uf g is surjective. Moreover, by hypothesis there is
some such g such that r(p) = n for all p ∈ X(g). Replacing f by f g we may assume
that r(p) = n for all p ∈ X(f ). For all such p, up : Rpn → Mp is therefore a sur-
jective endomorphism from a rank n free module to itself. Since finitely generated
modules are Hopfian, up is an isomorphism. By the local nature of isomorphisms
(Proposition 7.14) we conclude uf is an isomorphism, so Mf is free.
(ii) =⇒ (i): By Theorem 7.29, M is Z-locally free: there exists a finite Z-family
{fi }Q
i∈I such that for all i ∈ I, Mfi is finitely generated and free. Thus the mod-
n
Qn i=1 Mfi is finitely generated and projective over the faithfully flat R-algebra
ule
i=1 Rfi , so by faithfully flat descent (Theorem 3.111) M itself is projective. 

Corollary 13.36. Let R be a ring with Spec R irreducible (e.g. a domain).


For a finitely generated R-module M , the following are equivalent:
(i) R is projective.
(ii) R is Z-locally free.
(iii) R is locally free.
(iv) R is flat.
Exercise 13.35. Prove Corollary 13.36.

11. The Forster-Swan Theorem


For a prime ideal p of a ring R, put k(p) = Rp /pRp . For a fnitely generated
R-module M , we let µ(M ) denote the minimal cardinality of a set of R-module
11. THE FORSTER-SWAN THEOREM 255

generators for M . Recall that if (R, m) is local, then by Nakayama’s Lemma we


have
µ(M ) = dimR/m M/mM.
For p ∈ Spec R let µp (M ) be the minimal cardinality of a set of generators for the
Rp -module Mp .
Lemma 13.37. Let M be a finitely generated R-module, and let S be a finite
subset of supp M . Then there is m ∈ M such that for all p ∈ S, the image of M in
Mp /pMp is nonzero, and thus
∀p ∈ S, µp (M/hmi) = µp (M ) − 1.
Proof. We go by induction on the number of elements s of S. The base case
s = 0 is clear, so let s ≥ 1 and suppose the statement holds when #S = s − 1.
We may order the elements of S as p1 , . . . , ps so as to have p1 · · · ps−1 6⊂ ps ; choose
xs ∈ p1 · · · ps−1 \ ps . By induction, there is m1 ∈ M such that for all 1 ≤ i ≤ s − 1,
the image of m1 in Mpi /pi Mpi is nonzero. We are done if the image of m1 in
Mps /ps Mps is nonzero, so assume otherwise. Let m2 ∈ M have nonzero image in
Mps /ps Mps . Then m1 +xs m2 has nonzero image in Mpi /pi Mpi for all 1 ≤ i ≤ s. 
Theorem 13.38. (Forster-Swan [Fo64], [Sw67]) Let R be a ring such that the
space MaxSpec R is Noetherian, and let M be a finitely generated R-module. Then
(32) µ(M ) ≤ sup (µp (M ) + dim JSpec R/p) .
p∈JSpec R

Proof. We follow an exposition of Swan’s proof due to C.-L. Chai.


We may assume that supp∈JSpec R dim JSpec R/p < ℵ0 , for otherwise the con-
clusion is trivial. Since µp (M ) ≤ µ(M ) for all p ∈ Spec R, we have
µFS (M ) := sup (µp (M ) + dim JSpec R/p) < ℵ0 .
p∈JSpec R

We go by induction on µFS . If µF S (M ) = 0 then µm (M ) = 0 for all m ∈ MaxSpec R,


so M = 0 and thus µ(M ) = 0.
Induction Step: Suppose µFS (M ) ≥ 1 and suppose the result holds for finitely
generated modules R-modules N with µFS (N ) < µFS (M ). Let
S = {p ∈ JSpec R | µp (M ) + dim JSpec R/p = µFS (M )}.
For n ∈ N, put
Xn (M ) = {p ∈ JSpec R | µp (M ) ≥ n}.
Theorem 13.34 implies that Xn (M ) is closed in JSpec R. For q ∈ S, put n = µq (M )
and Z = {q}, so Z ⊂ Xn (M ). We claim that Z is an irreducible component of
Xn (M ): if not, there is p ∈ Xn (M ) such that p ( q and
µp (M ) + dim JSpec R/p > µq (M ) + dim JSpec R/q = µFS (M ),
contradiction. Because MaxSpec R is Noetherian, so is JSpec R and thus also
Xn (M ). It follows (JSpec R is sober, hence Kolmogorov) that S is finite. By
Lemma 13.37, there is m ∈ M such that µq (M/hmi) = µq (M ) − 1 for all q ∈ S.
Thus for all p ∈ JSpec R we have
µFS (M/hmi) ≤ k − 1,
so µ(M/hmi) ≤ k − 1 by induction. It follows that µ(M ) ≤ k. 
256 13. Spec R AS A TOPOLOGICAL SPACE

Corollary 13.39. Let M be a finitely generated module over a semilocal ring.


a) We have µ(M ) = supm∈MaxSpec R µm (M ).
b) If M is projective, then M is free iff the rank function rM is constant.
Proof. a) Since MaxSpec R is finite and separated, it is Noetherian and dis-
crete. So dim JSpec R = 0 and JSpec R = MaxSpec R. Apply Forster-Swan.
b) Free modules have constant rank. Conversely, suppose rM (m) = n for all
m ∈ MaxSpec R. By part a), µ(M ) = n, so M is free by Proposition 7.19. 
CHAPTER 14

Integral Extensions

1. First properties of integral extensions


If S is a ring extension of R – i.e., R ⊂ S – we will say that an element α of S is
integral over R if there exist a0 , . . . , an−1 ∈ R such that
αn + an−1 αn−1 + . . . + a1 α + a0 = 0.
Note that every element α ∈ R satisfies the monic polynomial t − α = 0, so is
integral over R.
Theorem 14.1. Let R ⊂ T be an inclusion of rings, and α ∈ T . the following
are equivalent:
(i) α is integral over R.
(ii) R[α] is finitely generated as an R-module.
(iii) There exists an intermediate ring R ⊂ S ⊂ T such that α ∈ S and S is finitely
generated as an R-module.
(iv) There exists a faithful R[α]-submodule M of T that is finitely generated as an
R-module.
Proof. (i) =⇒ (ii): If α is integral over R, there exist a0 , . . . , an−1 ∈ R such
that
αn + an−1 αn−1 + . . . + a1 α + a0 = 0,
or equivalently
αn = −an−1 αn−1 − . . . − a1 α − a0 .
This relation allows us to rewrite any element of R[α] as a polynomial of degree at
most n − 1, so that 1, α, . . . , αn−1 generates R[α] as an R-module.
(ii) =⇒ (iii): Take T = R[α].
(iii) =⇒ (iv): Take M = S.
(iv) =⇒ (i): Let m1 , . . . , mn be a finite set of generators for M over R, and express
each of the elements mi α in terms of these generators:
n
X
αmi = rij mj , rij ∈ R.
j=1

Let A be the n × n matrix αIn − (rij ); then recall from linear algebra that
AA∗ = det(A) · In ,
where A∗ is the “adjugate” matrix (of cofactors). If m = (m1 , . . . , mn ) (the row
vector), then the above equation implies 0 = mA = mAA∗ = m det(A) · In . The
latter matrix equation amounts to mi det(A) = 0 for all i. Thus • det(A) = •0 on
M , and by faithfulness this means det(A) = 0. Since so that α is a root of the
monic polynomial det(T · In − (aij )). 
257
258 14. INTEGRAL EXTENSIONS

Exercise 14.1. Let S be a finitely generated R-algebra. Show that the following
are equivalent:
(i) S/R is integral.
(ii) S is finite over R (as an R-module!).
In particular, if S/R is an extension and α1 , . . . , αn are all integral over R, then
R[α1 , . . . , αn ] is a finitely generated R-module.
Proposition 14.2. (Integrality is preserved under quotients and localizations)
Let S/R be an integral ring extension.
a) Let J be an ideal of S. Then S/J is an integral extension of R/(J ∩ R).
b) Let T be a multiplicatively closed subset of nonzero elements of R. Then ST is
an integral extension of RT .
Proof. a) First note that the kernel of the composite map R ,→ S → S/J is
J ∩ R, so that R/(J ∩ R) ,→ S/J is indeed a ring extension. Any element of S/J is
of the form x + J for x ∈ S, and if P (t)tn + an−1 tn−1 + . . . + a1 t + a0 = 0 ∈ R[t] is
a polynomial satisfied by x, then reducing coefficientwise gives a monic polynomial
P (t) ∈ R/(J ∩ R) satisfied by x.
b) Let J = {s ∈ S | ∃t ∈ T | ts = 0}, an ideal of S. Let T be the image of T
in R/(J ∩ R). Then ST ∼ = (S/J)T and JT ∼ = (R/(J ∩ R))T , so we may assume that
the maps R → RT and S → ST are injective. Let xy ∈ ST with x ∈ S, y ∈ T . Let
P (t) = tn + an−1 tn−1 + . . . + a0 ∈ R[t] be a monic polynomial satisfied by x. Then
 n  n−1
x an−1 x a0
+ + . . . + n = 0,
y y y y
x
showing that y is integral over RT . 

Lemma 14.3. Let R ⊂ S ⊂ T be an inclusion of rings. If α ∈ T is integral over


R, then it is also integral over S.
Proof. If α is integral over R, there exists a monic polynomial P ∈ R[t] such
that P (α) = 0. But P is also a monic polynomial in S[t] such that P (α) = 0, so α
is also integral over S. 
Lemma 14.4. Let R ⊂ S ⊂ T be rings. If S is a finitely generated R-module
and T is a finitely generated S-module, then T is a finitely generated R-module.
Proof. If α1 , . . . , αr generates S as an R-module and β1 , . . . , βs generates T
as an S-module, {αi βj }1≤i≤r,1≤j≤s generates T as an R-module: for α ∈ T ,
X XX
α= bj βj = (aij αi )βj ,
j i j

with bj ∈ S and aij ∈ R. 


Corollary 14.5. (Transitivity of integrality) If R ⊂ S ⊂ T are ring exten-
sions such that S/R and T /S are both integral, then T /R is integral.
Proof. For α ∈ T , let αn + bn−1 αn−1 + . . . + b1 α + b0 = 0 be an integral
dependence relation, with bi ∈ S. Thus R[b1 , . . . , bn−1 , α] is finitely generated over
R[b1 , . . . , bn−1 ]. Since S/R is integral, R[b1 , . . . , bn−1 ] is finite over R. By Lemma
14.4, R[b1 , . . . , bn−1 , α] is a subring of T containing α and finitely generated over
R, so by Theorem 14.1, α is integral over R. 
2. INTEGRAL CLOSURE OF DOMAINS 259

Corollary 14.6. If S/R is a ring extension, then the set IS (R) of elements
of S which are integral over R is a subring of S, the integral closure of R in S.
Thus R ⊂ IS (R) ⊂ S.
Proof. If α ∈ S is integral over R, R[α1 ] is a finitely generated R-module.
If α2 is integral over R it is also integral over R[α1 ], so that R[α1 ][α2 ] is finitely
generated as an R[α1 ]-module. By Lemmma 14.4, this implies that R[α1 , α2 ] is a
finitely generated R-module containing α1 ± α2 and α1 · α2 . By Theorem 14.1, this
implies that α1 ± α2 and α1 α2 are integral over R. 

If R ⊂ S such that IS (R) = R, we say R is integrally closed in S.


Proposition 14.7. Let S be a ring. The operator R 7→ IS (R) on subrings of
R is a closure operator in the abstract sense, namely it satisfies:
(CL1) R ⊂ IS (R),
(CL2) R1 ⊂ R2 =⇒ IS (R1 ) ⊂ IS (R2 ).
(CL3) IS (IS (R)) = IS (R).
Proof. (CL1) is the (trivial) Remark 1.1. (CL2) is obvious: evidently if R1 ⊂
R2 , then every element of S which satisfies a monic polynomial with R1 -coefficients
also satisfies a monic polynomial with R2 -coefficients. Finally, suppose that α ∈ S
is such that αn + an−1 αn−1 + . . . + a1 α + a0 = 0 for ai ∈ IS (R). Then each ai
is integral over R, so R[a1 , . . . , an ] is finitely generated as an R-module, and since
R[a1 , . . . , an , α] is finitely generated as an R[a1 , . . . , an ]-module, applying Lemma
14.4 again, we deduce that α lies in the finitely generated R-module R[a1 , . . . , an , α]
and hence by Theorem 14.1 is integral over R. 

2. Integral closure of domains


Until further notice we restrict to the case in which R ⊂ S are domains.
Proposition 14.8. Let R ⊂ S be an integral extension of domains.
a) R is a field iff S is a field.
b) An extension of fields is integral iff it is algebraic.
Proof. a) Suppose first that R is a field, and let 0 6= α ∈ S. Since α is
integral over R, R[α] is finitely generated as an R-module, and it is well-known in
field theory that this implies R[α] = R(α). Indeed, taking the polynomial of least
degree satisfied by α, say α(αn−1 + an−1 αn−2 + . . . + a1 ) = −a0 , then 0 6= a0 ∈ R
is invertible, so
−(αn−1 + an−1 αn−2 + . . . + a1 ) 1
= ,
a0 α
and S is a field. Conversely, if S is a field and a ∈ R, then R[a−1 ] is finite-
dimensional over R, i.e., there exist ai ∈ R such that
a−n = an−1 a−n+1 + . . . + a1 a−1 + a0 .
Multiplying through by an−1 gives
a−1 = an−1 + an−2 a + . . . + a1 an−2 + a0 an−1 ∈ R,
completing the proof of part a). Over a field every polynomial relation can be
rescaled to give a monic polynomial relation, whence part b). 
260 14. INTEGRAL EXTENSIONS

Remark: A more sophisticated way of expressing Proposition 14.8 is that if S/R is


an integral extension of domains, then dim R = 0 iff dim S = 0. Later we will see
that in fact dim R = dim S under the same hypotheses.

If R ⊂ S are fields, IS (R) is called the algebraic closure of R in S.


Exercise 14.2. a) Let S/R be an extension of fields. If S is algebraically
closed, then so is IS (R).
b) Deduce: if R = Q, S = C, then IS (R) is an algebraically closed, algebraic
extension of Q, denoted Q and called the field of all algebraic numbers.
Theorem 14.9. Let S/R be an extension of domains, and let T ⊂ R be a
multiplicatively closed subset. Then IT −1 S (T −1 R) = T −1 IS (R). In other words,
localization commutes with integral closure.
Proof. Let K be the fraction field of R and L the fraction field of S. Then
T −1 IS (R) is the subring of L generated by T −1 and the elements of S which are inte-
gral over R. Since both of these kinds of elements of T −1S are integral over T −1 R
and integral elements form a subring, we must have T −1 IS (R) ⊂ IT −1 S (T −1 R).
Conversely, let x ∈ T −1 S be integral over T −1 (R), so there are b0 , . . . , bn−1 ∈
T −1 (R) such that
xn + bn−1 xn−1 + . . . + b1 x + b0 = 0.
We may take a common denominator t ∈ T such that x = st and for all 0 ≤ i ≤ n−1,
bi = ati . Making this substitution and multiplying through by tn , we get
sn + an−1 sn−1 + tan−2 sn−2 + . . . + tn−2 a1 s + tn−1 = 0.
Thus s is integral ove R and x = s
t ∈ T −1 IS (R). 
Proposition 14.10. Let S/R be an extension of domains. Let K be the fraction
field of R and M the fraction field of S. Then the fraction field of IS (R) is IM (K).
Proof. We write L for the fraction field of IS (R). First we show IM (K) ⊂ L:
let x ∈ IS (K)• , so there are a0 , . . . , an−1 ∈ K such that xn + an−1 xn−1 + . . . +
a1 x + a0 = 0. After clearing denominators and relabelling, we get a0 , . . . , an ∈ R
such that
an xn + an−1 xn−1 + . . . + a1 x + a0 = 0.
Multiplying through by an−1n , we get
(an x)n + an−1 (an x)n−1 + . . . + a1 an−2
n (an x) + ann−1 a0 = 0,
which shows an x ∈ IS (R), so x ∈ IS (R) · (R \ 0)−1 ⊂ L.
The reverse inclusion L ⊂ IM (K) is quite similar, since an arbitrary nonzero
element x of L is of the form α β with α, β integral over R. But then certainly α and
β are both integral over K – i.e., algebraic over K, and then so also are β −1 and
x= α n
β . So again x satisfies a polynomial an t + . . . with coefficients in R and then
an x is integral over R, hence an x and an are algebraic over K and thus x = anxx is
algebraic over K so lies in IM (K). 
Example: If R = Z, S = C, then IS (R) is called the ring of all algebraic
integers, and often denoted Z. By Proposition 14.10, its fraction field is the field
Q of all algebraic numbers. This turns out to be a very interesting ring, and it will
crop up several times in the sequel as an example or counterexample. For instance:
3. SPECTRAL PROPERTIES OF INTEGRAL EXTENSIONS 261

Exercise 14.3. Show: Z is not finitely generated as a Z-module.1


Corollary 14.11. Let R be a domain with fraction field K and M/K an
arbitrary field extension. Put S = IM (R). Then S is integrally closed.
Proof. By Proposition 14.10, the field of fractions L of S is the algebraic
closure of K in M . If x ∈ L is integral over R, then since L ⊂ M it lies in
IM (R) = S. 

Let R be a domain with fraction field K. We say that R is integrally closed


if IK (R) = R, i.e., if any element of the fraction field satisfying a monic integral
polynomial with R-coefficients already belongs to R. It follows immediately from
Proposition 14.7 that in any case IK (R) is integrally closed.

Exercise 14.4. Let R = Z[ −3] = Z[t]/(t2 + 3). Show that R is not integrally
closed, and compute its integral closure.
The geometric terminology for an integrally closed domain is normal. The process
of replacing R by its integral closure IK (R) is often called normalization.

3. Spectral properties of integral extensions


Going down (GD): If we have I1 ⊃ I2 of R and J1 ∈ Spec S such that J1 ∩ R = I1 ,
there exists J2 ∈ Spec S such that J2 ⊂ J1 and J2 ∩ R = I2 .
Lemma 14.12. Let R be a local ring with maximal ideal p and S/R an integral
extension. Then the pushed forward ideal pS is proper.
P
Proof. Suppose not: then there exist pi ∈ p, si ∈ S such that 1 = i si pi .
Therefore any counterxample would take place already in the finite R-module
R[s1 , . . . , sd ]. By induction on d, it is enough to consider the case of n = 1:
S = R[s]. Consider as usual a relation
(33) sn = an−1 sn−1 + . . . + a1 s + a0 , ai ∈ R
of minimal possible degree n. If 1 ∈ pS then we have
(34) 1 = p0 + p1 s + . . . + pk sk , pi ∈ p.
In view of (33) we may assume k ≤ n − 1. Since 1 − p0 is not in the maximal ideal
of the local ring R, it is therefore a unit; we may therefore divide (34) by 1 − p0
and get an equation of the form
1 = p01 s + . . . + p0q sq , p0i ∈ p.
This shows that s ∈ S × . Replacing a0 = a0 · 1 in (33) by a0 (p01 s + . . . + p0q sq ), we
get an integral dependence relation which is a polynomial in s with no constant
term. Since s is a unit, we may divide through by it and get an integral dependence
relation of smaller degree, contradiction. 

Theorem 14.13. An integral ring extension S/R satisfies property (LO):2 every
prime ideal p of R is of the form S ∩ P for a prime ideal P of S.

1In fact it is not even a Noetherian ring, so not even finitely generated as a Z-algebra.
2Or, lying over.
262 14. INTEGRAL EXTENSIONS

Proof. For p a prime ideal of R, we denote – as usual – by Rp the localization


of R at the multiplicatively closed subset R \ p. Then Rp is local with unique
maximal ideal pRp , and if we can show that there exists a prime ideal Q of Sp lying
over pRp , then the pullback P = Q ∩ S to S is a prime ideal of S lying over p. By
Lemma 14.12, there exists a maximal ideal Q ⊃ pS and then Q ∩ R is a proper
ideal containing the maximal ideal p and therefore equal to it. 
Corollary 14.14. (Going up theorem of Cohen-Seidenberg [CS46]) Let S/R
be an integral extension and p ⊂ q be two prime ideals of R. Let P be a prime ideal
of S lying over p (which necessarily exists by Theorem 14.13). Then there exists a
prime ideal Q of S containing P and lying over q.
Proof. Apply Theorem 14.13 with R = R/p S = S/P and p = q/p. 
Corollary 14.15. (Incomparability) Suppose S/R is integral and P ⊂ Q are
two primes of S. Then P ∩ R 6= Q ∩ R.
Proof. By passage to S/P, we may assume that P = 0 and S is a domain,
and our task is to show that any nonzero prime ideal P of S lies over a nonzero
ideal of R. Indeed, let 0 6= x ∈ P, and let P (t) = tn + an−1 tn−1 + . . . + a0 ∈ R[t]
be a monic polynomial satisfied by x; we may assume a0 6= 0 (otherwise divide by
t). Then a0 ∈ xS ∩ R ⊂ P ∩ R. 
Corollary 14.16. Let S/R be an integral extension, P a prime ideal of S
lying over p. Then P is maximal iff p is maximal.
Proof. First proof: Consider the integral extension S/P/(R/p); we want
to show that S/P is a field iff R/p is a field. This is precisely Proposition 14.8a).
Second proof: If p is not maximal, it is properly contained in some maximal
ideal q. By the Going Up Theorem, there exists a prime Q ⊃ P lying over q, so
P is not maximal. Conversely, suppose that p is maximal but P is not, so there
exists Q ) P. Then Q ∩ R is a proper ideal containing the maximal ideal p, so
Q ∩ R = p = P ∩ R, contradicting the Incomparability Theorem. 
Invoking Going Up and Incomparability to (re)prove the elementary Corollary 14.16
is overkill, but these more sophisticated tools also prove the following
Corollary 14.17. Let S/R be an integral extension of rings. Then the Krull
dimensions of R and S are equal.
Proof. Suppose p0 ( p1 ( . . . ( pd are primes in R. Applying Theorem 14.1,
we get a prime P0 of S lying over p0 , and then repeated application of the Going Up
Theorem yields a chain of primes P0 ( P1 ( . . . ( Pd , so that dim(S) ≥ dim(R).
Similarly, if we have a chain of prime ideals P0 ( . . . ( Pd of length d in S, then
Theorem 14.15 implies that for all 0 ≤ i < d, Pi ∩ R ( Pi+1 . 

4. Integrally closed domains


Let R ⊂ S be domains. Immediately from the definition of integrality, there is a
concrete way to show that x ∈ S is integral over R: it suffices to exhibit a monic
polynomial P ∈ R[t] with P (x) = 0. What if we want to show that x ∈ S is
not integral over R? It would suffice to show that R[x] is not a finitely generated
R-module, but exactly how to do this is not clear.
4. INTEGRALLY CLOSED DOMAINS 263


As an example, it is obvious that √
α = 2 is an algebraic integer, but unfortu-
nately it is not obvious that β = 22 is not an algebraic integer. (And of course we

need to be careful, because e.g. γ = 1+2 5 is an algebraic integer, since it satisfies
t2 + t − 1 = 0.) One thing to notice is that unlike α and γ, the minimal polynomial
of β, t2 − 12 , does not have Z-coefficients. According to the next result, this is
enough to know that β is not integral over Z.
Theorem 14.18. Let R be a domain with fraction field K, let S/R be an
extension domain, and x ∈ S an integral element over R.
a) Let P (t) ∈ K[t] be the minimal polynomial of x over K. Then P (t) ∈ IK (R)[t].
b) If R is integrally closed, the minimal polynomial of x has R-coefficients.
Proof. a) Let g ∈ R[t] be a monic polynomial satisfied by x. Let K be the
fraction field of R, and let P ∈ K[t] be the minimal polynomial of x over K,
and let M/K be the splitting field of P , so there are α1 , . . . , αd ∈ M such that
Qd
P = i=1 (t − αi ). There is h ∈ K[t] such that g = P h in K[t]. For all 1 ≤ i ≤ d
we have g(αi ) = 0, so each αi is integral over R. Since the roots of P lie among
the αi ’s they are also integral over R, hence so too are the coefficients of P , being
polynomial expressions in the roots.
b) This follows immediately from part a). 
Exercise 14.5. Let R be a domain with fraction field K. Let S/R be an
extension such that for every x ∈ S which is integral over R, the minimal polynomial
P (t) ∈ K[t] has R-coefficients. Show: R is integrally closed.
Theorem 14.19. (Local nature of integral closure) For a domain R, the fol-
lowing are equivalent:
(i) R is integrally closed.
(ii) For all prime ideals p of R, Rp is integrally closed.
(iii) For all maximal ideals m of R, Rm is integrally closed.
Proof. Let K be the fraction field of R. Assume (i), and let p ∈ Spec R. By
Theorem 14.9, the integral closure of Rp in K is Rp . Evidently (ii) =⇒ (iii).
Assume (iii), and let x be an element of K which is integral over R. Then for every
maximal idealTm of R, certainly x is integral overTRm , so by assumption x ∈ Rm
and thus x ∈ m Rm . By Corollary 7.15 we have m Rm = R. 
Exercise 14.6. Let R be an integrally closed domain with fraction field K, L/K
an algebraic field extension, S the integral closure of R in L and G = Aut(L/K).
a) Show: for every σ ∈ G, σ(S) = S.
b) For P ∈ Spec S and σ ∈ G, show σ(P) = {σ(x) | x ∈ P} is a prime ideal of S.
c) Show: P ∩ R = σ(P) ∩ R.
In conclusion, for every p ∈ Spec R, there is a well-defined action of G on the
(nonempty!) set of prime ideals P of S lying over p.
Lemma 14.20. Let R be a domain with fraction field K of characteristic p > 0,
let L/K be a purely inseparable algebraic extension of K (possibly of infinite degree),
and let S be the integral closure of R in L. For any p ∈ Spec R, rad(pR) is the
unique prime of S lying over p.
Exercise 14.7. Prove Lemma 14.20.
(Suggestions: recall that since L/K is purely inseparable, for every x ∈ L, there
264 14. INTEGRAL EXTENSIONS

a
exists a ∈ N such that xp ∈ K. First observe that rad(pR) contains every prime
ideal of S which lies over p and then show that rad(pR) is itself a prime ideal.)
Theorem 14.21. (Going Down Theorem of Cohen-Seidenberg [CS46]) Let R be
an integrally closed domain with fraction field K, and let S be an integral extension
of R. If p1 ⊂ p2 are prime ideals of R and P2 is a prime ideal of S lying over p2 ,
then there exists a prime ideal P1 of S which is contained in P2 and lies over p1 .
Proof. Let L be a normal extension of K containing S, and let T be the
integral closure of R in L. In particular T is integral over S, so we may choose
Q2 ∈ Spec T lying over P2 and also Q1 ∈ Spec T lying over p1 . By the Going Up
Theorem there exists Q0 ∈ Spec T containing Q1 and lying over p2 . Both Q2 and Q0
lie over p2 , so by Theorem 14.36 there exists σ ∈ Aut(L/K) such that σ(Q0 ) = Q2 .
Thus σ(Q1 ) ⊂ σ(Q0 ) = Q2 and σ(Q1 ) lies over p1 , so that setting P1 = σ(Q1 ) ∩ S
we have P1 ∩ R = p1 and P1 ⊂ σ(Q0 ) ∩ S = Q2 ∩ S = P2 . 
Remark: In [AM, Chapter 5] one finds a proof of Theorem 14.21 which avoids all
Galois-theoretic considerations. However it is significantly longer than the given
proofs of Theorems 14.36 and 14.21 combined and – to me at least – rather opaque.

5. The Noether Normalization Theorem


5.1. The classic version.
Theorem 14.22. (Noether Normalization) Let k be a field, and let R be a
domain with fraction field K. Suppose that R is moreover a finitely generated k-
algebra, generated say by elements x1 , . . . , xm . Then:
a) There exists d ∈ Z, 0 ≤ d ≤ m, and algebraically independent elements y1 , . . . , yd ∈
R such that R is finitely generated as a module over the polynomial ring k[y1 , . . . , yd ]
– or equivalently, that R/k[y1 , . . . , yd ] is an integral extension.
b) We have dim R = d = trdeg(K/k) (the transcendence degree of K/k).
Proof. a) (Jacobson) The result is trivial if m = d, so we may suppose m > d.
Then the yi are algebraically dependent over k: there exists a nonzero polynomial
X
f (s1 , . . . , sm ) = aJ sj11 · · · sjmm , aJ ∈ k[s1 , . . . , sm ]

with f (x1 , . . . , xm ) = 0. Let X be the set of monomials sJ = sj11 · · · sjmm occuring


in f with nonzero coefficients. To each such monomial we associate the univariate
polynomial
j1 + j2 t + . . . + jm tm−1 ∈ Z[t].
The polynomials obtained in this way from the elements of X are distinct. Since
a univariate polynomial over a field has only finitely many zeroes, it follows that
there exists a ≥ 0 such that the integers j1 + j2 a + . . . + jm am−1 obtained from the
monomials in X are distinct. Now consider the polynomial
m−1
f (s1 , sa1 + t1 , . . . , sa1 + tm ) ∈ k[s, t].
We have
m−1 X m−1
f (s1 , sa1 + t1 , . . . , s1a + tm ) = aJ sj11 (sd1 + t2 )j2 · · · (sa1 + tm )jm
J
X m−1
= aJ sj11 +j2 a+...+jm a + g(s1 , t2 , . . . , ym ),
J
5. THE NOETHER NORMALIZATION THEOREM 265

m m−1
in which the degree of g in s1 is less than that of J aJ sj11 +j2 a+...+j a
P
. Hence
× a am−1
for suitable β ∈ k , βf (s1 , s1 + t2 , . . . , s1 + tm ) is a monic polynomial in x1
i−1
with k[t2 , . . . , tm ]-coefficients. Putting wi = xi − xa1 for 2 ≤ i ≤ m, we get
m−1
βf (x1 , xd1 + w2 , . . . , xa1 + wm ) = 0,

so that x1 is integral over R0 = k[w2 , . . . , wm ]. By induction on the number of gener-


ators, R0 has a transcendence base {yi }di=1 such that R0 is integral over k[y1 , . . . , yd ].
Thus R is integral over k[y1 , . . . , yd ] by transitivity of integrality.
b) Since R/k[y1 , . . . , yd ] is integral, by Corollary 14.17 the Krull dimension of R is
equal to the Krull dimension of k[y1 , . . . , yd ], which by Corollary 12.17 is d. Since
R is finitely generated as a k[y1 , . . . , yd ] algebra, by Proposition 14.10 K is finitely
generated as a k(y1 , . . . , yd )-module, so trdeg K/k = trdeg k(y1 , . . . , yd ) = d. 

5.2. Separable Noether Normalization.

Let K be a field and let K an algebraic closure of K. A field extension L/K


is regular if L ⊗K K is a field (equivalently, a domain).
For a field extension L/K, we say K is algebraically closed in L if any ele-
ment of L which is algebraic over K lies in K. It is an easy exercise to show that if
L/K is regular, K is algebraically closed in L. The converse is true in characteristic
zero, but in positive characteristic we need a further hypothesis:

Theorem 14.23. Let L/K be a field extension.


a) The following are equivalent:
(i) L/K is regular.
(ii) L/K is separable and K is algebraically closed in L.
b) In particular, if K is perfect and algebraically closed in L, then L/K is regular.

Proof. a) The key result here is Mac Lane’s Theorem [FT, §12]: a field
−∞
extension L/K is separable iff L and K p are linearly disjoint over K.
(i) =⇒ (ii): If L/K is regular, then K is algebraically closed in L. Further, L and
−∞
K are linearly disjoint over K, hence L and K p are linearly disjoint over K.
−∞
0
(ii) =⇒ (i): Let K = K p
and L = L⊗K K . Since L⊗K K = (L⊗K K 0 )⊗K 0 K =
0 0

L ⊗K K, it is enough to show that L0 is a field and L0 ⊗K 0 K is a field. Now L0 is a


0 0

field by Mac Lane’s Theorem, and since K 0 is perfect, K/K 0 is a Galois extension,
and thus by [FT, §12.3], since L0 ∩ K = K 0 , L0 and K are linearly disjoint over K 0 .
b) If K is perfect, every extension of K is separable. Apply part a). 

Theorem 14.24. (Separable Noether Normalization) Let k be a field, and let


R be a domain that is finitely generated as a k-algebra. Assume moreover that the
fraction field L of R is a regular extension of k.
a) There exists d ∈ Z, 0 ≤ d ≤ m, and algebraically independent elements y1 , . . . , yd ∈
R such that R is finitely generated as a k[y1 , . . . , yd ]-module and L/k(y1 , . . . , yd ) is
a finite separable field extension.
b) The integer d is equal to both the Krull dimension of R and the transcendence
degree of K/k.

For now we refer the reader to [Ei, Cor. 16.18] for the proof.
266 14. INTEGRAL EXTENSIONS

5.3. Noether normalization over a domain.


Theorem 14.25. (Noether Normalization II) Let R ⊂ S be domains with
S finitely generated as an R-algebra. There exists a ∈ R• and y1 , . . . , yd ∈ S
algebraically independent over the fraction field of R such that Sa (the localization
of S at the multiplicative subset generated by a) is finitely generated as a module
over T = Ra [y1 , . . . , yd ].
Proof. (K.M. Sampath) Let K be the fraction field of R and let x1 , . . . , xm
be a set of R-algebra generators for S. Then
S 0 := S ⊗R K = K[x1 , . . . , xm ]
is finitely generated over K (as above, the xi ’s need not be algebraically inde-
pendent). Applying Theorem 14.22, we get algebraically independent elements
y1 , . . . , yd ∈ S 0 such that S 0 is a finitely generated T 0 := K[y1 , . . . , yd ]-module.
Multiplying by a suitable element of R× , we may assume yi ∈ S for all i.
Since S 0 is finitely generated as a T 0 -module, it is integral over T 0 . For
1 ≤ i ≤ m, xi satisfies a monic polynomial equation with coefficients in T 0 :
y1n + Pi,1 (y1 , . . . , yd )yin−1 + . . . + Pi,n = 0.
Let a be the product of the denominators of all coefficients of all the polynomials
Pi,k . It follows that Sa is integral and finitely generated as a T = Ra [y1 , . . . , yd ]-
algebra, hence it is finitely generated as a T -module. 
Exercise 14.8. In the setting of Theorem 14.25 suppose S is a graded R-
algebra. Show: we may take all the yi to be homogeneous elements.
5.4. Applications.

The Noether Normalization Theorem is one of the foundational results in algebraic


geometry: geometrically, it says that every integral affine variety of dimension d is
a finite covering of affine d-space Ad . Thus it allows us to study arbitrary varieties
in terms of rational varieties via branched covering maps. It is almost as important
as a theorem of pure algebra, as even the “soft” part of the result, that the Krull
dimension of an integral affine k-algebra is equal to the transcendence degree of its
fraction field, is basic and useful.

One of the traditional applications of Noether Normalization is to prove Hilbert’s


Nullstellensatz. As we have seen, it is fruitful to channel proofs of the Nullstellen-
satz through Zariski’s Lemma, and this is no exception.
Proposition 14.26. Noether Normalization implies Zariski’s Lemma.
Exercise 14.9. Prove Proposition 14.26.
Theorem 14.27. Let k be a field, and let R be a domain which is finitely
generated as a k-algebra, with fraction field K. Then:
a) The Krull dimension of R is equal to the transcendence degree of K/k.
b) Every maximal chain of prime ideals in R has length dim R.
Proof. a) By Noether normalization, R is finite over k[t1 , . . . , td ], so K is
finite over k(t1 , . . . , td ). Thus the transcendence degree of K/k is d. On the other
hand, R is integral over k[t1 , . . . , td ] so by Theorem 14.17 dim R = dim k[t1 , . . . , td ].
6. SOME CLASSICAL INVARIANT THEORY 267

By Corollary 12.17, dim k[t1 , . . . , td ] = d.


b) See Madapusi-Sampath, p. 119... 

6. Some Classical Invariant Theory


Let R be a commutative ring, let G be a finite group, and suppose G acts on R by
automorphisms, i.e., we have a homomorphism ρ : G → Aut(R). We define
RG = {x ∈ R |∀g ∈ G, gx = x},
the ring of G-invariants – it is indeed a subring of R.

Remark: As in the case of rings acting on commutative groups, we say that G-


action on R is faithful if the induced homomorphism ρ : G → Aut(R) is injective.
Any G-action induces a faithful action of G/ ker(ρ), so it is no real loss of general-
ity to restrict to faithful G-actions. We will do so when convenient and in such a
situation identify G with its isomorphic image in Aut R.

The simplest case is that in which R = K is a field. Then K G is again a field


and K/K G is a finite Galois extension. Conversely, for any finite Galois extension
K/F , F = K Aut(K/F ) . This characterization of Galois extensions was used by
E. Artin as the foundation for an especially elegant development of Galois theory
(which swiftly became the standard one). Note also the analogy to topology: we
have the notion of a finite Galois covering Y → X of topological spaces as one for
which the group G = Aut(Y /X) of deck transformations acts freely and properly
discontinuously on Y such that Y /G = X.

The branch of mathematics that deals with invariant rings under linear group ac-
tions is called classical invariant theory. Historically it was developed along
with basic commutative algebra and basic algebraic geometry in the early 20th
century, particularly by Hilbert. Especially, Hilbert’s work on the finite generation
of invariant rings was tied up with his work on the Basis Theorem.

σ(a). Then NG (a) ∈ RG , so we have a map


Q
For a ∈ R, put NG (a) = σ∈G

NG : R → R G .
Note that NG is not a homomorphism of additive groups. However, when R is a
domain, there is an induced map
NG : R• → (RG )•
which is a homomorphism of monoids, so induces a homomorphism on unit groups.
Exercise 14.10. Let R[t] be the univariate polynomial ring over R. Show:
there is a unique action of G by automorphisms of G on R[t] extending the G-
action on R and such that gt = t. Show that (R[t])G = RG [t].
Proposition 14.28. For a finite group G acting on R, R/RG is integral.
Proof. For x ∈ R, define
Y
Φx (t) = NG (t − x) = (t − gx),
g∈G
268 14. INTEGRAL EXTENSIONS

so Φx (t) ∈ (R[t])G = RG [t]. Thus Φx (t) is a monic polynomial with RG -coefficients


which is satisfied by x. 
Base extension: Suppose that G is a finite group acting faithfully on R. Moreover,
let A be a ring and f : A → R be a ring homomorphism, so R is an A-algebra.
Suppose moreover that f (A) ⊂ RG . In such a situation we say that G acts on R
by A-automorphisms and write G ⊂ Aut(R/A).

Suppose we have another A-algebra A0 . We can define an action of G on R ⊗A A0


by putting g(x ⊗ y) := gx ⊗ y. We say that the G-action is extended to A0 .
Proposition 14.29. In the above setup, suppose that A0 is a flat A-algebra.
Then there is a natural isomorphism

RG ⊗A A0 → (R ⊗A A0 )G .
Proof. Madapusi p. 65-66. 
Corollary 14.30. Let G be a finite group acting on the ring R, and let S ⊂ RG
be a multiplicatively closed set. Then (S −1 R)G = S −1 RG .
Exercise 14.11. Prove Corollary 14.30.
In particular, suppose R is a domain with fraction field K, and let F be the fraction
field of RG . Then the G-action on R extends to a G-action on K, and Corollary
14.30 gives K G = F . Thus the invariant theory of domains is compatible with the
Galois theory of the fraction fields.
Proposition 14.31. If R is integrally closed, so is RG .
Proof. Let x ∈ K be integral over RG . Then x is also integral over R, and
since R is integally closed in L we have x ∈ R. Thus x ∈ R∩K = R∩LG = RG . 

Theorem 14.32. (Noether [No26]) Suppose that R is a finitely generated


algebra over some field k with k = k G . Then:
a) R is finitely generated as an RG -module.
b) RG is a finitely generated k-algebra.
Proof. a) Since R is a finitely generated k-algebra and k ⊂ RG , R is a finitely
generated RG -algebra. But by Proposition 14.28 R/RG is integral. So R is finitely
generated as an RG -module.
b) By part a), the Artin-Tate Lemma (Theorem 8.52) applies to the tower of rings
k ⊂ RG ⊂ R. The conclusion is as desired: RG is a finitely generated k-algebra. 
Remark: The title of [No26] mentions “characteristic p”. In fact, when k has char-
acteristic 0 the result had been proven by Hilbert significantly earlier [Hi90], and
moreover for certain actions of infinite linear groups, like SLn (k). But Noether’s
formulation and proof give an excellent illustration of the economy and power of
the commutative algebraic perspective.

Let us make contact with the setup of classical invariant theory: let k be a field,
V a finite-dimensional vector space and ρ : G → Autk (V ) a linear representation
of G on V . Let k[V ] = Sym(V ∨ ) be the algebra of polynomial functions on V . If
we choose a k-basis e1 , . . . , en of V and let x1 , . . . , xn be the dual basis of V ∨ , then
6. SOME CLASSICAL INVARIANT THEORY 269

k[V ] = k[x1 , . . . , xn ] is a polynomial ring in n independent indeterminates. There


is an induced action of G on k[V ], namely for f ∈ k[V ] we put (gf )(x) = f (g −1 x).

All of our above results apply in this situation. Especially, Theorem 14.32 ap-
plies to tell us that the ring k[V ]G is finitely generated as a k-algebra, or a finite
system of invariants. Of course, we did not so much as crease our sleeves, let
alone roll them up, to establish this: for a concretely given finite group G and action
on a k-vector space V , it is of interest to explicitly compute such a finite system.
Moreover, the polynomial ring k[V ] is integrally closed: in the next section we will
see that it is a unique factorization domain and that this is a stronger property.
Therefore Proposition 14.31 applies to show that k[V ]G is integrally closed. This
is actually quite a robust and useful procedure for producing integrally closed rings.

Example: Let k be a field, n ∈ Z+ , let V = k n , G = Sn be the symmetric group,


and let G act on V by permuting the standard basis elements e1 , . . . , en . We will
compute k[V ]G . Namely, for 1 ≤ i ≤ n, we define the ith elementary symmetric
function si (t1 , . . . , tn ) as follows: let X be an independent indeterminate and put
n
Y n
X
f (X) = (X − ti ) = X n + (−1)i si (t1 , . . . , tn )X n−i .
i=1 i=1

Theorem 14.33. The invariant ring k[V ]Sn is a polynomial k-algebra on the
elementary symmetric functions s1 , . . . , sn .
Proof. Step 1: Explicitly, we have
s1 = t1 + . . . + tn ,
X
s2 = ti tj ;
i<j

each si is the sum of all nk monomials of degree k. Clearly k[s1 , . . . , sn ] ⊂ k[V ]Sn .


Step 2: For any finite group G of automorphisms of a field L, L/K G is a Galois


extension with Aut(L/LG ) = G. Take L = k(V ) and note that k(V ) is the splitting
field of the separable polynomial f ∈ k(s1 , . . . , sn )[x], so k(V )G = k(s1 , . . . , sn ).
Step 3: Because k(t1 , . . . , tn )/k(s1 , . . . , sn ) is a finite extension, the transcendence
degree of k(s1 , . . . , sn )/k is equal to the transcendence degree of k(t1 , . . . , tn )/k,
namely n. It follows that the elements s1 , . . . , sn are algebraically independent,
i.e., k[s1 , . . . , sn ] is a polynomial ring.
Step 4: As in the proof of Proposition 14.31,
k[t1 , . . . , tn ]Sn = k[t1 , . . . , tn ] ∩ k(s1 , . . . , sn ) = k[s1 , . . . , sn ]. 

The above example is well-known and extremely useful, but gives a misleadingly
simple impression of classical invariant theory. One can ask how often the ring
of invariants of a finite group action on a polynomial ring is again a polynomial
ring, and there is a nice answer to this. But let’s back up a step and go back to
“rational invariant theory”: if G acts on k[x1 , . . . , xn ], then as above it also acts on
the fraction field k(x1 , . . . , xn ) and we know that k(x1 , . . . , xn )/k(x1 , . . . , xn )G is a
finite Galois extension. But must k(x1 , . . . , xn )G itself be a rational function field,
as it was in the example above? This is known as Noether’s Problem: it was
first posed by E. Noether in 1913. It is natural and important, for an affirmative
270 14. INTEGRAL EXTENSIONS

answer would allow us to realize every finite group as a Galois group (i.e., the auto-
morphism group of a Galois extension) of Q thanks to a famous theorem of Hilbert.
For more than half of the twentieth century, Noether’s problem remained open.
Finally, in 1969 R.G. Swan (yes, the same Swan as before!) found a representation
of the cyclic group of order 47 on a finite-dimensional Q-vector space for which the
invariant field is not a rational function field [Sw69]. Too bad – this was arguably
the best shot that anyone has ever taken at the Inverse Galois Problem over Q.3

Example: Let k be a field of characteristic different from 2, let V = k 2 , and


consider the action of the two-element group G = {±1} on V by −1 acting as the
scalar matrix −1. The induced action on k[V ] = k[x, y] takes x 7→ −x and y 7→ −y.
This is, apparently, a not very interesting representation of a not very interesting
group. But the invariant theory is very interesting!
Exercise 14.12. a) Show: k[V ]G is generated as a k-algebra by x2 , y 2
and xy.
b) Show: k[V ]G is isomorphic to the k-algebra k[A, B, C]/(AB − C 2 ).
c) Show: k[V ]G is not isomorphic to k[x, y].4
d) Show that nevertheless the fraction field of k[V ]G is rational, i.e., is iso-
morphic to k(X, Y ) for independent indeterminates X and Y .
Before signing off on our quick glimpse of classical invariant theory, we cannot resist
mentioning one more classic theorem in the subject. It answers the question: when
is the invariant subalgebra k[V ]G isomorphic to a polynomial algebra over k?

Let ρ : G ,→ GL(V ) be a faithful representation of G on a finite-dimensional


k-vector space V . An element g ∈ GL(V ) is a pseudoreflection if it has finite
order and pointwise fixes a hyperplane W in V . (Equivalently, a pseudoreflection
has characteristic polynomial (t − 1)dim V −1 (t − ζ), where ζ is a root of unity in k.)
Exercise 14.13. If k is formally real, any nontrivial pseudoreflection has order
2 – i.e., it really is a hyperplane reflection.
A faithful representation ρ of G is a pseudoreflection representation of G if
ρ(G) is generated by pseudorflections.
Theorem 14.34. (Shephard-Todd-Chevalley-Serre) Let k be a field, and let
ρ : G ,→ GL(V ) be a faithful finite-dimensional k-linear representation.
a) If k[V ]G is a polynomial algebra, then ρ is a pseudoreflection representation.
b) If ρ is a pseudoreflection representation and char k - #G, then k[V ]G is a poly-
nomial algebra.
Proof. See [Be, §7.2]. 
In the modular case char k | #G, there are pseudoreflection representations for
which k[V ]G is not a polynomial algebra. However, work of Kemper and Malle
[KM99] shows that even in the modular case, if ρ is an irreducible pseudoreflection
representation then the invariant field k(V )G is purely transcendental over k.
3Actually, Serre’s Topics in Galois Theory describes a conjecture of J.-L. Colliot-Thélène –
roughly a weaker form of Noether’s problem – which would still imply that every finite group is
a Galois group over Q. I am not aware of any progress on this conjecture.
4Suggestion: the proof of Theorem 15.42 shows that C[A, B, C]/(AB − C 2 ) is not a unique
factorization domain. The argument goes through with C replaced by k.
7. GALOIS EXTENSIONS OF INTEGRALLY CLOSED DOMAINS 271

Exercise 14.14. It follows from Theorem 14.34 the fundamental theorem on


symmetric functions that the standard permutation representation of the symmetric
group Sn on k n is a pseudoreflection representation. Show this directly.

7. Galois extensions of integrally closed domains


Proposition 14.35. Let G be a finite group acting by automorphisms on a
ring R, with invariant subring RG . Let ι : RG ,→ R, and let p ∈ Spec RG .
a) There is a natural action of G on the fiber (ι∗ )−1 (p) – i.e., on the set of primes
P of R such that ι∗ P = p.
b) The G-action on the fiber (ι∗ )−1 (p) is transitive.
Proof. Let P ∈ Spec R and σ ∈ G. Define
σP = {σx | x ∈ P}.
It is straightforward to verify that σP is a prime ideal of R (if you like, this follows
from the fact that Spec is a functor). Moreover (σP) ∩ RG is the set of all elements
σx with x ∈ P such that for all g ∈ G, gσx = σx. As g runs through all elements
of G, so does gσ −1 , hence (σP) ∩ RG = P ∩ RG = p.
b) Let P1 , P2 be two primes of R lying over a prime p of RG . Let x ∈ P1 . Then
NG (x) ∈ P1 ∩ RG = p ⊂ P2 . Since S P2 is prime, there exists at least one σ ∈ G
such that σx ∈ P2 , and thus P1 ⊂ σ∈G σP2 . By Prime Avoidance (Lemma 8.47),
there exists σ ∈ G such that P1 ⊂ σP2 . Since R/RG is integral, Incomparability
(Corollary 14.15) yields P1 = σP2 . 
Theorem 14.36. Let R be an integrally closed domain with fraction field K,
let L/K be a normal algebraic field extension (possibly of infinite degree), and let S
be the integral closure of R in L. Let p ∈ Spec R, and let Xp be the set of all prime
ideals of S lying over p. Then G = Aut(L/K) acts transitively on Xp .
Proof. Step 1: Suppose [L : K] = n < ∞, and write G = {σ1 = 1, . . . , σr }.5
Seeking a contradiction, suppose there are P1 , P2 ∈ Xp such that P2 6= σj−1 P1 for
all j. By Corollary 14.15, P2 is not contained in any σj−1 P1 , so by Prime Avoidance
(Lemma 8.47) there is x ∈ P2 \ j σj−1 P1 . Let q be the inseparable degree of L/K
S
Q q
and put y = j σj (x) . Thus y = NL/K (x), so y ∈ K. Moreover y is integral
over R, so y ∈ R. Since σ1 = 1, y ∈ P2 , so y ∈ P2 ∩ R = p ⊂ P1 , and thus, since
P1 is prime, σj (x) ∈ P1 for some j: contradiction!
Step 2: We will reduce to the case in which L/K is a Galois extension. Let G =
Aut(L/K) and K 0 = LG , so that L/K 0 is Galois and K 0 /K is purely inseparable.
Let R0 be the integral closure of R in K 0 . Then by Lemma 14.20 Spec R0 → Spec R
is a bijection. So we may as well assume that K 0 = K and L/K is Galois.
Step 3: For each finite Galois subextension M of L/K, consider the subset
F (M ) := {σ ∈ G | σ(P1 ∩ M ) = P2 ∩ M }.
Observe that F (M ) is a union of cosets of Gal(L/M ) hence is (open and) closed
in theQKrull topology. By Step 1, we have F (M ) 6= ∅. Moreover, the compositum
M = i Mi of any finite number {M T i } of finite Galois subextenions is again a finite
Galois subextension, and we have i F (Mi ) ⊃ F (M ) 6= ∅. Therefore as Mi ranges
through all finite Galois subextensions of L/K, {F (Mi )}i∈I is a family of closed
5We are assuming that L/K is normal, so L/K is separable iff L/K is Galois iff r = n.
272 14. INTEGRAL EXTENSIONS

subsets of the compact spaceTG satisfying the finite intersection condition, and it
follows that there exists σ ∈ i F (Mi ) = F (L) i.e., σ ∈ G such that σP1 = P2 . 

8. Almost Integral Extensions


We come now to a technical variant of the notion of integrality. This variant will
not be used until §19.4 on divisorial ideals. We honestly recommend that the reader
skip past this section for now and return only when the concept of complete integral
closure is needed and used.

Let R ⊂ S be rings. An element x ∈ S is almost integral over R if there is


a finitely generated R-submodule of S which contains xn for all n ∈ Z+ . We say
that S is almost integral over R if every element of S is almost integral over R.
Proposition 14.37. Let R ⊂ S be rings, and let x ∈ S.
a) If x is integral over R, it is almost integral over R.
b) If R is Noetherian and x is almost integral over R, then x is integral over R.
Proof. Let M = hR, xi. By Theorem 14.1, x is integral over R iff M is a
finitely generated R-module.
a) If x is integral over R, then M is a finitely generated R-submodule of S containing
xn for all n ∈ Z+ , so x is almost integral over R.
b) Suppose x is almost integral over R: there is a finitely generated R-submodule
N of S containing xn for all n ∈ Z+ . Then M ⊂ N , and since R is Noetherian and
N is finitely generated, M is finitely generated and x is integral over R. 
Remark: As the proof shows, an equivalent – and perhaps more perspicuous – way
of expressing the almost integrality condition is that, while integrality of x means
that M = hR, xiR is a finitely generated submodule of S, almost integrality means
that there is some finitely generated R-submodule of S containing M .

For rings R ⊂ S, the complete integral closure of R in S is the set of all


elements of S which are almost integral over R. A domain R is completely inte-
grally closed if its complete integral closure in its fraction field is R itself.
Theorem 14.38. Let R be a domain with fraction field K. The complete inte-
gral closure of R is the set of all x ∈ K such that there is r ∈ R• with rxn ∈ R for
all n ∈ Z+ .
Proof. Let x ∈ K be almost integral over R. Then there are h1 , . . . , hs ∈ K
such that R[x] ⊂ hh1 , . . . , hs iR . If r is the product of the denominators of the hi ,
then rxn ∈ R for all n ∈ Z+ .
Let x ∈ K be such that there is r ∈ R• with rxn ∈ R or all n ∈ Z+ . Then
R[x] ⊂ r−1 R. 
Tournant dangereux: If R ⊂ S and R0 is the complete integral closure of R in S,
then R0 is integrally closed in S [LM, Prop. 4.18]. However, the complete integral
closure of a domain in its fraction field need not be completely integrally closed
[LM, Exc. IV.14]! In other words, complete integral closure is unfortunately not a
closure operator on the set of subrings of a field in the sense of §2.1.
CHAPTER 15

Factorization

Let R be a domain, and x a nonzero, nonunit element of R. We say that x is


irreducible if for any y, z ∈ R such that x = yz, one of y or z is a unit.

For any unit u ∈ R× , we get factorizations of the form x = u · (u−1 x), so ev-
ery x has at least these factorizations, which we wish to regard as “trivial”. On the
other hand, y and z cannot both be units, for then x would also be a unit. Let us
then define a factorization of a nonzero nonunit a ∈ R as a product
a = x1 · · · xn ,
such that each xi is irreducible. We say that two factorizations
a = x1 · · · xn = y1 · · · ym
are equivalent if the multisets of associated principal ideals {{(xi )}} = {{(yj )}}
are equal. More concretely, this means that m = n and that there is a bijection
σ : {1, . . . , n} → {1, . . . , m} such that (yσ(i) ) = (xi ) for all 1 ≤ i ≤ n.

If factorizations always exist and any two factorizations of a given element are
equivalent, we say R is a unique factorization domain (UFD).

1. Kaplansky’s Theorem (II)


A basic and important result that ought to get covered at the undergraduate level
is that PID implies UFD. In fact this is easy to prove. What is more difficult is to
get a sense of exactly how UFDs are a more general class of rings than PIDs. In
this regard, an elegant theorem of Kaplansky seems enlightening.
Exercise 15.1. Let x be an element of a domain which can be expressed as
x = p1 · · · pn ,
such that for 1 ≤ i ≤ n, pi = (pi ) is a prime ideal. If then there exist principal
prime ideals q1 , . . . , qm such that (x) = q1 · · · qm , then m = n and there exists a
permutation σ of the integers from 1 to n such that qi = pσ(i) for all i.
Exercise 15.2. Let R be a domain, and let S be the set of all nonzero elements
x in R such that (x) can be expressed as a product of principal prime ideals. Show:
S is a saturated multiplicatively closed subset.
Theorem 15.1. (Kaplansky) a domain is a UFD iff every nonzero prime ideal
in R contains a prime element.
Proof. Suppose R is a UFD and 0 6= p ∈ Spec R. Let x ∈ p• , and write
x = p1 · · · pr
273
274 15. FACTORIZATION

a product of prime elements. Then x ∈ p implies pi ∈ p for some i, so (pi ) ⊂ p.


Conversely, assume each nonzero prime ideal of R contains a principal prime.
Let S be the set of all products of prime elements, so that by Exercise 15.2, S is
a saturated multiplicative subset. By Exercise 15.1, it is enough to show that S
contains all nonzero nonunits of R. Suppose for a contradiction that there exists
a nonzero nonunit x ∈ R \ S. The saturation of S implies S ∩ (x) = ∅, and then
by Theorem 5.24 there is a prime ideal p containing x and disjoint from S. But by
hypothesis, p contains a prime element p, contradicting its disjointness from S. 
Corollary 15.2. Let R be a domain.
a) If R is a UFD, then every height one prime ideal is principal. If R is Noetherian
and every height one prime ideal is principal, then R is a UFD.
b) If every ideal of R is principal, R is a UFD.
c) Conversely, if R is a UFD of dimension one, every ideal of R is principal.
Proof. We begin by recalling that in a domain R the height one primes are
the nonzero prime ideals p such that there is no prime ideal q with (0) ( q ( p.
a) If R is a UFD and p is a height one prime ideal, then by Theorem 15.1 p has
a prime element (π). Then (π) ⊂ p is a containment of prime ideals, and since p
has height one we have p = (π) is principal. Conversely, suppose R is Noetherian
and every height one prime ideal is principal. Let x ∈ R be irreducible: i.e., x is
a nonzero nonunit and x = yz implies y ∈ R× or z ∈ R× . Let p be a minimal
prime over (x). By Krull’s Hauptidealsatz, p has height one and so by assumption
p = (π). Thus π | x and since x is irreducible we must have (π) = (x). Thus we
have shown that every irreducible element of R is prime. As we will see §3 (and as
the reader may know from “general algebra”), a Noetherian domain in which each
irreducible element is prime is a UFD.
b) If every ideal of R is principal, then every ideal of R is finitely generated. So R
is Noetherian and part a) applies.
c) In a UFD of dimension one, part a) implies that every prime ideal is principal.
By Theorem 4.35, every ideal is principal. 
One may ask: in a not necessarily Noetherian domain, if every height one prime is
principal, must R be a UFD? The following exercise addresses this.
Exercise 15.3. a) Let R be a domain which satisfies (DCC) on prime ideals.
Show: R is a UFD iff every height one prime is principal.
b) Show: a domain R in which nonzero primes have infinite height yields a negative
answer to the question: every height one prime is principal but R is not a UFD.
c) Show: there are domains in which every nonzero prime ideal has infinite height.
(This is difficult. For the answer, see [AKP98, Ex. 2.2].)

2. Atomic domains, ACCP


A domain in which every nonzero nonunit can be factored into irreducibles is an
atomic domain.
Exercise 15.4. Show: the ring Z of all algebraic integers is not an atomic
domain. Indeed, since for every algebraic integer x, there exists an algebraic integer
y such that y 2 = x, there are no irreducible elements in Z!
The condition of factorization into irreducibles (in at least one way) holds in every
Noetherian domain. In fact, a much weaker condition than Noetherianity suffices:
2. ATOMIC DOMAINS, ACCP 275

Proposition 15.3. Let R be a domain in which every ascending chain of prin-


cipal ideals stabilizes. Then every nonzero nonunit factors into a product of irre-
ducible elements. In particular, a Noetherian domain is atomic.
Proof. Let R be a domain satisfying the ascending chain condition for prin-
cipal ideals (ACCP for short), and suppose for a contradiction that R is not an
atomic domain. Then the set of principal ideals generated by unfactorable ele-
ments is nonempty, so by our assumption there exists a maximal such element, say
I = (a). Evidently a is not irreducible, so we can begin to factor a: a = xy where
x and y are nonunits. But this means precisely that both principal ideals (x) and
(y) properly contain (a), so that by the assumed maximality of (a), we can factor
both x and y into irreducibles: x = x1 · · · xm , y = y1 · · · yn . But then
a = x1 · · · xm y1 · · · yn
is a factorization of a, contradiction. 

This proposition motivates us to consider also the class of domains which satisfy
the ascending chain condition for principal ideals (ACCP).
Exercise 15.5. Suppose R ,→ S is an extension of rings such that S × ∩ R =
R× . (In particular, this holds for integral extensions.) Show that S satisfies
(ACCP) implies R satisfies (ACCP). Does the converse hold?
We have just seen that (ACCP) implies atomicity. The proof shows that under
(ACCP) we can always obtain an expression of a given nonzero nonunit by a finite
sequence of “binary factorizations” i.e., replacing an element x with y1 · y2 , where
y1 and y2 are nonunits whose product is x. After a bit of thought, one is inclined to
worry that it may be possible that this factorization procedure fails but nevertheless
irreducible factorizations exist. This worry turns out to be justified:
Theorem 15.4. There exists an atomic domain which does not satisfy (ACCP).
Proof. See [Gr74]. 

However, the following strenghtening of atomicity does imply ACCP:

A domain R is a bounded factorization domain BFD if it is atomic and for


each nonzero nonunit a ∈ R, there exists a positive integer N (a) such that in any
irreducible factorization a = x1 · · · xr we have r ≤ N (a).
Proposition 15.5. A UFD is a BFD.
Proof. An immediate consequence of the definitions. 

Proposition 15.6. A BFD satisfies (ACCP).


Proof. Let R be a BFD. Suppose for a contradiction that (xi )i∈Z+ is a strictly
ascending chain of principal ideals. We therefore have
x0 = y1 x1 = y1 y2 x2 = . . . = y1 · · · yn xn = · · · ,
with each xi , yi a nonunit. Since R is atomic, we can refine each factorization into
an irreducible factorization, but clearly an irreducible refinement of y1 · · · yn xn has
at least n + 1 irreducible factors, contradicting BFD. 
276 15. FACTORIZATION

3. EL-domains
An element x of a domain R is prime if the principal ideal (x) is a prime ideal.
Equivalently, x satisfies Euclid’s Lemma: if x | yz, then x | y or x | z.
Proposition 15.7. A prime element is irreducible.
Proof. If x is reducible, then x = yz with neither y nor z a unit, so that
yz ∈ (x) but y ∈
6 (x), z 6∈ (x). 
However, it need not be the case that irreducible elements are prime!
√ √
Example: Let R = Z[ −5]. Then 2, 3 and 1 ± −5 are all irreducible, but
√ √
2 · 3 = (1 + −5)(1 − −5)
shows that none of them are prime.

Exercise 15.6. Check all √ these assertions.
√ Hint: Define
√ N (a + b −5)
√ =
a2 + 5b2 . Check that N ((a + b −5)(c + d −5)) = N (a + b −5)N (c + d −5).
Show that√ α | β (in R) =⇒ N (α) | N (β) (in Z), and use this to show that
2, 3, 1 ± −5 are irreducible but not prime.
It is tempting to call a domain in which all irreducible elements are prime “Eu-
clidean,” but this terminology is already taken for domains satisfying a generaliza-
tion of the Euclidean algorithm (c.f. §16.3). So we will, provisionally, call a ring in
which irreducible elements are prime an EL-domain. (EL = Euclid’s Lemma).
Theorem 15.8. For a domain R, the following are equivalent:
(i) R is a UFD.
(ii) R satisfies (ACCP) and is an EL-domain.
(iii) R is an atomic EL-domain.
Proof. i) =⇒ (ii): In the previous section we saw UFD =⇒ BFD =⇒
(ACCP). We show UFD implies EL-domain: let x ∈ R be irreducible and suppose
x | yz. Let y = y1 · · · ym and z = z1 · · · zn be irreducible factorizations of y and z.
Then the uniqueness of irreducible factorization means that x must be associate to
some yi or to some zj , and hence x | y or x | z: R is an EL-domain.
(ii) =⇒ (iii) follows immediately from Proposition 15.3.
(iii) =⇒ (i): This is nothing else than the usual deduction of the fundamental
theorem of arithmetic from Euclid’s Lemma: in a factorization domain we have at
least one irreducible factorization of a given nonzero nonunit x. If we also assume
irreducibles are prime, we may compare any two irreducible factorizations: suppose
x = y1 · · · ym = z1 · · · zn .
Then y1 is a prime element so divides zj for some j. WLOG, relabel to assume
j = 1. Since z1 is irreducible, we have y1 = u1 z1 and thus we may cancel to get
y2 · · · ym = (u−1
1 z2 )z3 · · · zn .

Continuing inQthis way we find that each yi is associate to some zj ; when we get
down to 1 = j zj we must have no factors of zj left, so m = n and R is a UFD. 
We can now deduce the following important result, a characterization of Noetherian
UFDs among all Noetherian domains.
4. GCD-DOMAINS 277

Theorem 15.9. For a Noetherian domain R, the following are equivalent:


(i) Every height one prime ideal of R is principal.
(ii) R is a UFD.
Proof. (i) =⇒ (ii): By Theorem 15.8, it is sufficient to prove that R is an
EL-domain, so let x ∈ R be irreducible. Let p be a minimal prime containing x. By
Krull’s Hauptidealsatz (Theorem 8.44), p has height one, so by assumption p = (p)
is principal. Thus x = up for some u ∈ R, and since x and p are both irreducible,
u ∈ R× , (x) = p, and x is a prime element.
(ii) =⇒ (i): This implication is a special case of Kaplansky’s Theorem 15.1 (and
thus holds without the Noetherian assumption on R). 

4. GCD-domains
For elements a and b of a domain R, a greatest common divisor is an element
d of R such that: d | a, d | b and for e ∈ R with e | a, e | b, e | d.
Exercise 15.7. Show: if d is a gcd of a and b, then an element d0 of R is a
gcd of a and b iff (d) = (d0 ). In particular, any two gcd’s are associate.
If a and b have a gcd, it would be more logically sound to write gcd(a, b) to mean
the unique principal ideal whose generators are the various gcd’s of a and b. It
is traditional however to use the notation gcd(a, b) to denote an element, with the
understanding that in general it is only well-defined up to multiplication by a unit.1

More generally, for elements a1 , . . . , an in a domain R, a greatest common divi-


sor is an element d of R such that d | ai for all i and if e | ai for all i then e | d.
If a GCD of (a1 , . . . , an ) exists, it is unique up to associates, and we denote it by
gcd(a1 , . . . , an ). As above, it can be characterized as the unique minimal princi-
pal ideal containing ha1 , . . . , an i. Moreover, these setwise GCDs can be reduced to
pairwise GCDs.
Exercise 15.8. Let a, b, c be elements of a domain R and assume that all
pairwise GCD’s exist in R. Then gcd(a, b, c) exists and we have gcd(a, gcd(b, c)) =
gcd(a, b, c) = gcd(gcd(a, b), c).
A domain R is a GCD-domain if for all a, b ∈ R, gcd(a, b) exists. By the above
remarks, it would be equivalent to require that gcd(a1 , . . . , an ) for all n-tuples of
elements in R.
Proposition 15.10. (GCD Identities) Let R be a GCD-domain. Then:
a) For all a, b, c ∈ R, gcd(ab, ac) = a gcd(b, c).
a b
b) For all a, b ∈ R \ {0}, gcd( gcd(a,b) , gcd(a,b) ) = 1.
c) For all a, b, c ∈ R, gcd(a, b) = gcd(a, c) = 1, then gcd(a, bc) = 1.
d) For all a, b, c ∈ R, gcd(a, b + ac) = gcd(a, b).
e) For all a, a1 , . . . , an , b1 , . . . , bn , c ∈ R, gcd(a, b1 +ca1 , . . . , bn +can ) = gcd(a, b1 , . . . , bn ).
Proof. a) Let x = gcd(ab, ac). Then a | ab and a | ac so a |x: say ay = x.
Since x | ab and x | ac, y | b and y | c, so y | gcd(b, c). If z | b and z | c, then az | ab
and az | ac, so az | x = ay and z | y. Therefore gcd(b, c) = y = a1 gcd(ab, ac).
1In some rings, principal ideals have canonical generators: e.g. in the integers we may take
the unique positive generator and in k[t] we may take the unique monic generator. Under these
circumstances, a common convention is to let gcd(a, b) stand for this canonical generator.
278 15. FACTORIZATION

b) This follows immediately from part a).


c) Suppose gcd(a, b) = gcd(a, c) = 1, and let t divide a and bc. Then t divides ab
and bc so t | gcd(ab, bc) = b gcd(a, c) = b. So t divides gcd(a, b) = 1.
d) If d divides both a and b, it divides both a and b + ac. If d divides both a and
b + ac, it divides b + ac − c(a) = b.
e) We have
gcd(a, b1 + ca1 , . . . , bn + can ) = gcd(a, gcd(a, b1 + ac1 ), . . . , gcd(a, bn + acn ))
= gcd(a, gcd(a, b1 ), . . . , gcd(a, bn )) = gcd(a, b1 , . . . , bn ). 
Proposition 15.11. A GCD-domain is an EL-domain.
Proof. This follows from: gcd(x, y) = gcd(x, z) = 1 =⇒ gcd(x, yz) = 1. 
Theorem 15.12. Consider the following conditions on a domain R:
(i) R is a UFD.
(ii) R is a GCD-domain.
(iii) R is an EL-domain: irreducible elements are prime.
a) We have (i) =⇒ (ii) =⇒ (iii).
b) If R is an ACCP-domain, (iii) =⇒ (i).
Proof. a) (i) =⇒ (ii): Let x, y be nonzero elements of R. We may write
x = f1 · · · fr g1 · · · gs , y = uf1 · · · fr h1 · · · ht ,
where the f ’s, g’s and h’s are prime elements, (gj ) 6= (hk ) for all j, k and u ∈ R× .
Then f1 · · · fr is a gcd for x and y.
(ii) =⇒ (iii): This is Proposition 15.11.
b) (iii) + (ACCP) =⇒ (i): This is Theorem 15.8. 
Corollary 15.13. For a Noetherian domain R, the following are equivalent:
(i) R is a UFD.
(ii) R is a GCD-domain.
(iii) R is an EL-domain.
We now present some simple results that are long overdue. An extremely useful
fact in algebra is that any UFD is integrally closed in its fraction field. We give a
slightly stronger result and then recall a classical application.
Theorem 15.14. A GCD-domain is integrally closed.
Proof. Let R be a GCD-domain with fraction field K. Suppose x ∈ K satisfies
xn + an−1 xn−1 + . . . + a1 x + a0 = 0, ai ∈ R.
Write x = st with s, t ∈ R, t 6= 0. By Proposition 15.10b), after dividing by the gcd
we may assume gcd(s, t) = 1. Plugging in x = st and clearing denominators gives
sn = − an−1 tsn−1 + . . . + a1 tn−1 s + a0 tn ,


so t | sn . But by Proposition 15.10c) gcd(sn , t) = 1, so t ∈ R× and x ∈ R. 


Corollary 15.15. An algebraic integer which is a rational number is an in-
teger: Z ∩ Q = Z.
Exercise 15.9. Prove Corollary 15.15.
5. GCDS VERSUS LCMS 279


Thus e.g. one can derive the irrationality of 2: it is a root of the monic polynomial
equation t2 − 2 = 0 but evidently not an integer, so cannot be rational.
Proposition 15.16. (Compatibility of GCD’s with localization) Let R be a
GCD-domain and S a multiplicative subset of R. Then:
a) The localization S −1 R is again a GCD-domain.
b) For all x, y ∈ R• , if d is a GCD for x and y in R, then it is also a GCD for x
and y in S −1 R.
Exercise 15.10. Prove Proposition 15.16.
Theorem 15.17. Let R be a UFD with 2 ∈ R× . Let f ∈ R be squarefree – i.e.,
not divisible by the square of any nonunit – and not a square. Then
p
S = R[ f ] := R[x]/hx2 − f i
is an integrally closed domain.
Proof. Let K be the fraction field of R. By Theorem 15.14 R is integrally
closed, and thus f is not a square in K – if so, it would also be a square in R –
so x2 − f is irreducible√in K[x] and thus L := K[x]/(x2 − f ) is a quadratic √ field
2
extension
√ of K. Write f for the element x + (x − f ) of L, so L = K( f ) and
S = R[ f ]. In particular, S is a domain. √
Let α be an element of L that is integral over R, so α = a + b f with a, b ∈ K.
If b = 0 then α ∈ K hence α ∈ R since R is integrally closed in K, so assume b 6= 0.
Then the minimal polynomial of α is
P (t) = t2 − 2at + (a2 − b2 f ).
By Theorem 14.18 we have −2a ∈ R and a2 − b2 f ∈ R. Since 2 ∈ R× , we get
that a ∈ R and then that b2 f ∈ R. Since f is squarefree, this implies that b ∈ R:
otherwise, ordp (b) ≤ −1 for some prime element p, so ordp (b2 f ) ≤ −2 + 1 < 0. So
α ∈ S. Thus S is integrally closed. 

Exercise √15.11. Show: in Theorem 15.17 the hypothesis 2 ∈ R× is necessary.


(E.g. show: Z[ 3] is not integrally closed.)

5. GCDs versus LCMs


The definition of GCDs in a domain has an evident analogue for least common mul-
tiples. Namely, if a and b are elements of a domain R, a least common multiple
of a and b is an element l such that for all m ∈ R with a | m and b | m then l | m.

Many of the properties of GCD’s carry over immediately to LCM’s. For instance,
if l is an LCM of a and b, then l0 ∈ R is an LCM of a and b iff l0 is associate to l.
Proposition 15.18. Let a and b be elements in a domain R. Then lcm(a, b)
exists iff the ideal (a) ∩ (b) is principal, in which case the set of all LCM’s of a and
b is the set of all generators of (a) ∩ (b).
Proof. This is straightforward and left to the reader. 

LCM’s exist in any UFD: if


a = xa1 1 · · · xar r , b = xb11 · · · xbrr ,
280 15. FACTORIZATION

with ai , bi ∈ N. Then
max(a ,b )
l = x1 1 1
· · · xmax(a
r
r ,br )

is a greatest common divisor of a and b. Now the simple identity


∀a, b ∈ N, min(a, b) + max(a, b) = a + b
implies that for a, b in any UFD R we have
gcd(a, b) lcm(a, b) ∼ ab.
This identity further suggests that the existence of either one of gcd(a, b), lcm(a, b)
implies the existence of the other. However, this turns out only to be half correct!
Theorem 15.19. For a, b in a domain R, the following are equivalent:
(i) lcm(a, b) exists.
(ii) For all r ∈ R \ {0}, gcd(ra, rb) exists.

Proof. Step 1: i) =⇒ (ii). Suppose that there exists a least common multiple
of a and b, say l. We claim that d := ab l is a greatest common divisor of a and b.
(Since ab is a common divisor of a and b, l | ab, so indeed d ∈ R.) Indeed, suppose
that e | a and e | b. Then since ab e is a common multiple of a and b, we must have
l | ab
e and this implies e | ab
l . Thus d is a GCD of a and b.
Step 2: Suppose that for r ∈ R \ {0} and a, b ∈ R, gcd(ra, rb) exists. Then we
claim that gcd(a, b) exists and gcd(ra, rb) = r gcd(a, b). Put g := gcd(ra,rb)
r , which
is clearly an element of D. Since gcd(ra, rb) divides ra and rb, g divides a and b.
Conversely, if e | a and e | b, then re | ra and re | rb so er | gcd(ra, rb) and e | g.
Step 3: We claim that if l := lcm(a, b) exists then so does lcm(ra, rb) for all
r ∈ R \ {0}. First observe that rl is a common multiple of ra and rb. Now
suppose m is a common multiple of ra and rb, say m = xra = yrb = r(xa − yb).
Thus r | m and a | m m m
r , b | r . So l | r and rl | m. Thus lcm(ra, rb) = r lcm(a, b).
Step 4: (ii) =⇒ (i). We may assume that a and b are nonzero, since the other
cases are trivial. Suppose gcd(ra, rb) exists for all r ∈ R \ {0}. We claim that
ab
l := gcd(a,b) is an LCM of a and b. Clearly l is a common multiple of a and b. Now
suppose that m is a common multiple of a and b. Then ab divides both ma and mb,
ab
so ab | gcd(ma, mb). By Step 2, gcd(ma, mb) = m gcd(a, b). Thus gcd(a,b) | m. 
Theorem 15.20. (Khurana, [Kh03, Thm. 4]) Let d ≥ 3 be an integer such
that d + 1 is not prime, and √ number p and k ≥ 2.
√ write d + 1 = pk for a prime
Then in the domain R = Z[ −d], the elements p and 1 + −d have a GCD but no
LCM.
Proof. Step 1: We claim that p is irreducible as an element of R. Indeed,
√ if
it were reducible, then by the multiplicativity of the norm map N (a + b −d) =
a2 + dp2 we could write it as p = αβ, with
p2 = N (p) = N (αβ) = N (α)N (β),
and, since α, β are nonunits, N (α), N (β) > 1. But then N (α) = N (β) = p, i.e.,
there would be a, b ∈ Z such that a2 + db2 = p. But this is not possible: either
ab = 0, in which the
√ left hand side is a perfect square,
√ or a2 + db2 ≥√d + 1 > p.
1 1
Step 2: gcd(p, 1 + −d) = 1. Indeed, since p + p −d 6∈ R, p - 1 + −d.

Step 3: We claim that kp and k(1 + −d) do not have a GCD. Indeed, by Step
2 of the proof of Theorem 15.19, if any GCD exists then k is a GCD. Then, since
6. POLYNOMIAL RINGS OVER UFDS 281

√ √ √ √ √
1 + −d divides both √ (1 − −d)(1 + −d) = 1 + d = kp and k(1 + −d), 1 + −d
divides gcd(kp, k(1 + −d) = k, i.e., there exist a, b ∈ Z such that
√ √ √
k = (1 + −d)(a + b −d) = (a − db) + (a + b) −d,
i.e., a = −b and k = a − db = a + da = a(1 + d) and d + 1 | k, contradicting the
fact that 1 < k < d + 1. √
Step 4: It follows from Theorem 15.19 that lcm(p, 1 + −d) does not exist. 
Khurana produces similar√examples even when d + 1 is prime, which implies that
for no d ≥ 3 is Rd = Z[ −d] a GCD-domain. (In fact, since (Rd , +) ∼= Z2 , Rd
is an abstract number ring and hence Noetherian, so the notions of EL-domain,
GCD-domain and UFD are all equivalent.) Let us give an independent proof:

Theorem 15.21. For no d ≥ 3 is Rd = Z[ −d] an EL-domain.
Proof. As in the proof of Theorem 15.20 above, the easy observation that
the equation a2 + db2 = 2 has no integral solutions implies that the element 2 is
irreducible in Rd . Now, since (quite trivially)
√ −d is
√ a square modulo 2, there exists
x ∈ Z such that 2 | x2 + d = (x + −d)(x − −d). But now, if Rd were an
EL-domain, the irreducible element √ 2 would be prime √ and hence Euclid’s Lemma
would apply to show that 2 | x ± −d, i.e., that x2 + 12 −d ∈ Rd , which is a clear
contradiction ( 12 is not an integer!). 
Theorem 15.19 has the following immediate consequence:
Corollary 15.22. (Cohn, [?, Thm. 2.1]) For a domain R, the following are
equivalent:
(i) Any two elements of R have a greatest common divisor.
(ii) Any two elements of R have a least common multiple.
Thus we need not define an “LCM-domain”: these are precisely the GCD domains.

6. Polynomial rings over UFDs


Our goal in this section to show that if R is a UFD, then a polynomial ring in any
number (possibly infinite) of indeterminates is again a UFD. This result generalizes
a familiar fact from undergraduate algebra: if k is a field, k[t] is a UFD. The cor-
responding fact that polynomials in k[t1 , . . . , tn ] factor uniquely into irreducibles is
equally basic and important, and arguably underemphasized at the pre-graduate
level (including high school, where factorizations of polynomials in at least two
variables certainly do arise).

If we can establish that R a UFD implies R[t] a UFD, then an evident induc-
tion argument using R[t1 , . . . , tn , tn+1 ] = R[t1 , . . . , tn ][tn+1 ] gives us the result for
polynomials in finitely many indeterminates over a UFD. It is then straightforward
to deduce the case for an arbitrary set of indeterminates.

There are several ways to prove the univariate case. Probably the most famous
is via Gauss’s Lemma. For this we need some preliminary terminology.

Let R be a domain, and consider a nonzero polynomial


f = an tn + . . . + a1 t + a0 ∈ R[t].
282 15. FACTORIZATION

We say f is primitive if x ∈ R, x | ai for all i implies x ∈ R× . In a GCD-


domain, this is equivalent to gcd(a1 , . . . , an ) = 1. In a PID, this is equivalent
to ha0 , . . . , an i = R. For a general domain, this latter condition is considerably
stronger: e.g. the polynomial xt + y ∈ k[x, y][t] is primitive but the coefficients do
not generate the unit ideal. Let us call this latter – usually too strong condition –
naively primitive.
Proposition 15.23. Let R be a domain, and f, g ∈ R[t] be naively primitive.
Then f g is naively primitive.
Proof. Suppose f and g are naively primitive but f g is not. Then by definition
the ideal generated by the coeffiicents of f is proper, so lies in some maximal ideal
m of R. For g ∈ R[t], write g for its image in the quotient ring (R/m)[t]. Then our
assumptions give precisely that f , g 6= 0 but f g = f g = 0. Thus f and g are zero
divisors in the domain (R/m)[t], a contradiction. 
If R is a GCD-domain and 0 6= f ∈ R[t], we can define the content c(f ) of f to be
the gcd of the coefficients of f , well-determined up to a unit. Thus a polynomial is
primitive iff c(f ) = 1.
Exercise 15.12. Let R be a GCD-domain and 0 6= f ∈ R[t].
a) Show: f factors as c(f )f1 , where f1 is primitive.
b) Let 0 6= a ∈ R. Show: c(af ) = ac(f ).
Theorem 15.24. (Gauss’s Lemma) Let R be a GCD-domain. If f, g ∈ R[t] are
nonzero polynomials, we have c(f g) = c(f )c(g).
If we assume the stronger hypothesis that R is a UFD, we can give a very trans-
parent proof along the lines of that of Proposition 15.23 above. Since this special
case may be sufficient for the needs of many readers, we will give this simpler proof
first, followed by the proof in the general case.
Proof. (Classical proof for UFDs) The factorization f = c(f )f1 of Exercise
15.10 reduces us to the following special case: if f and g are primitive, then so is
f g. Suppose that f g is not primitive, i.e., there exists a nonzero nonunit x which
divides all of the coefficients of f g. Since R is a UFD, we may choose a prime
element π | x. Now we may argue exactly as in the proof of Proposition 15.23:
(R/(π)[t] is a domain, f and g are nonzero, but f g = f g = 0, a contradiction. 
The proof of the general case uses the GCD identities of Proposition 15.10.
Proof. (Haible) As above, we may assume that f = an tn + . . . + a1 t + a0 , g =
m
bm t + . . . + b1 t + b0 ∈ R[t] are both primitive, and we wish to show that f g =
cm+n tm+n + . . . + c1 t + c0 is primitive. We go by induction on n. Since a primitive
polynomial of degree 0 is simply a unit in R, the cases m = 0 and n = 0 are both
trivial; therefore the base case m + n = 0 is doubly so. So assume m, n > 0. By
Proposition 15.10, we have
c(f g) = gcd(cn+m , . . . , c0 ) =
gcd(an bm , gcd(cn+m−1 , . . . , c0 )) | gcd(an , gcd(cn+m−1 , . . . , c0 ))·gcd(bm , gcd(cn+m−1 , . . . , c0 )).
Now
gcd(an , gcd(cn+m−1 , . . . , c0 )) = gcd(an , cn+m−1 , . . . , c0 )
= gcd(an , cn+m−1 − an bm−1 , . . . , cn − an b0 , cn−1 , . . . , c0 )
6. POLYNOMIAL RINGS OVER UFDS 283

= gcd(an , c((f − an tn )g).


Our induction hypothesis gives c((f − an tn )g) = c(f − an tn )c(g) = c(f − an tn ), so
gcd(an , cn+m−1 −an bm−1 , . . . , cn −an b0 , cn−1 , . . . , c0 ) = gcd(an , c(f −an tn )) = c(f ) = 1.
Similarly we have gcd(bm , gcd(cn+m−1 , . . . , c0 )) = 1, so c(f g) = 1. 

Corollary 15.25. Let R be a GCD-domain with fraction field K, and let


f ∈ R[t] be a polynomial of positive degree.
a) The following are equivalent:
(i) f is irreducible in R[t].
(ii) f is primitive and irreducible in K[t].
b) The following are equivalent:
(i) f is reducible in K[t].
(ii) There exist g, h ∈ R[t] such that deg(g), deg(h) < deg(f ) and f = gh.
Proof. a) Assume (i). Clearly an imprimitive polynomial in R[t] would be
reducible in R[t], so f irreducible implies c(f ) = 1. Suppose f factors nontrivially
in K[t], as f = gh, where both g, h ∈ K[t] and have smaller degree than f . By
Exercise 15.10, we may write g = c(g)g1 , h = c(h)h1 , with g1 , h1 primitive, and
then f = c(g)c(h)g1 h1 . But then g1 , h1 , being primitive, lie in R[t], and c(f ) =
c(g)c(h) = c(gh) ∈ R, so the factorization takes place over R[t], contradiction. (ii)
=⇒ (i) is similar but much simpler and left to the reader.
b) That (ii) =⇒ (i) is obvious, so assume (i). Because we can factor out the
content, it is no loss of generality to assume that f is primitive. Let f = g1 h1 with
g1 , h1 ∈ K[t] and deg(g1 ), deg(h1 ) < deg(f ). Because R is a GCD-domain, we may
write g = dg̃1 , h = dh̃2 with g̃, h̃ ∈ R[t] primitive. Then we have d1 d2 f = g̃ h̃, and
equating contents gives (d1 d2 ) = (1), so d1 , d2 ∈ R× and thus the factorization
f = gh has the properties we seek. 

We now give Gauss’s proof that a univariate polynomial ring over a UFD is a UFD.
Theorem 15.26. If R is a UFD, so is R[t].
Proof. Let K be the fraction field of R, and let f ∈ R[t]• . We know that K[t]
is a PID hence a UFD, so we get a factorization
f = cg1 · · · gr ,
with c ∈ R and each gi ∈ R[t] is primitive and irreducible. Then factoring c into
irreducibles gives an irreducible factorization of f . If we had another irreducible
factorization f = dh1 · · · hs , then unique factorization in K[t] gives that we have
r = s and after permuting the factors have gi = ui hi for all i, where ui ∈ K × .
Since both gi and hi are primitive, we must have ui ∈ R× , whence the uniqueness
of the factorization. 

This proof relies on knowing that K[t] is a UFD, which of course follows from
the fact that polynomial division gives a Euclidean algorithm, as one learns in an
undergraduate course. This is of course an adaptation of the proof that the ring Z
is a UFD (the Fundamental Theorem of Arithmetic) essentially due to Euclid.
It is interesting to find alternate routes to such basic and important results.
284 15. FACTORIZATION

Theorem 15.27. Let R be a domain with fraction field K.


a) If R is an ACCP-domain, so is R[t].
b) If R is a GCD-domain, so is R[t].
c) Thus, once again, if R is a UFD, so is R[t].
Proof. a) In an infinite ascending chain {(Pi )} of principal ideals of R[t],
deg Pi is a descending chain of non-negative integers, hence eventually constant.
Therefore for sufficiently large n we have Pn = an Pn+1 with an ∈ R and (an+1 ) ⊃
(an ). Since R is an ACCP domain, we have (an ) = (an+1 ) for sufficiently large n,
hence also (Pn ) = (Pn+1 ) for sufficiently large n.
b) (Haible, [Ha94]) Let f, g ∈ R[t]. We may assume that f g 6= 0. As usual, write
f = c(f )f˜ and g = c(g)g̃. Since K[t] is a PID, may take the gcd of f˜ and g̃ in
K[t], say d.˜ The choice of d˜ is unique only up to an element of K × , so by choosing
the unit appropriately we may assume that d˜ lies in R[t] and is primitive. We put
˜
d = gcd(c(f ), c(g)))d.
Step 1: We claim that d˜ is a gcd of f˜ and g̃ in R[t]. Since d˜ | f in K[t], we
˜
may write fd˜ = ab q with a, b ∈ R \ {0} and q ∈ R[t] primitive. Since bf˜ = ad, ˜ we
˜ ˜ b × ˜ ˜
have (b) = c(bf ) = c(ad) = (a), i.e., a ∈ R and thus d | f in R[t]. Similarly
d˜ | g̃. Moreover, sice d˜ ∈ f˜K[t] + g̃K[t], there exist u, v ∈ R[t] and c ∈ R \ {0}
with cd˜ = uf˜ + vg̃. Suppose h ∈ R[t] divides both f˜ and g̃. Then h | cd, ˜ and
˜ cd a
c(h) | c(f ) = (1). Writing h = b q with q ∈ R[t] primitive, and equating contents
˜
in bcd˜ = ahq, we get (bc) = (a), hence hd = ab ˜
c q ∈ R[t], so h | d.
Step 2: We claim that d is a gcd of f and g in R[t]. Certainly we have
˜ | (c(f )f˜) = (f ),
(d) = (gcd(c(f ), c(g)d)
so d | f . Similarly d | g. Conversely, let h ∈ R[t] divide f and g. Write h = c(h)h̃
for h̃ ∈ R[t] primitive. From h | f it follows that c(h) | c(f ) and thus h̃ | f˜ .
Similarly h | f so h̃ | g̃. Thus c(h) | gcd(c(f ), c(g)), h̃ | d˜ and thus finally h | d.
c) If R is a GCD domain and an ACCP domain, it is also an atomic EL domain,
hence a UFD by Theorem 15.8. 
Lindemann [Li33] and Zermelo [Ze34] (independently) gave (similar) striking proofs
of the Fundamental Theorem of Arithmetic avoiding all lemmas and packaging the
Euclidean division into a single inductive argument. Later several authors have
recorded analogous proofs of Gauss’s Theorem (Theorem 15.26): the earliest in-
stance we are aware of in the literature is due to S. Borofsky [Bo50]. We give a
third, “lemmaless” proof of Theorem 15.26 here.
Proof. It suffices to show that R[t] is an ACCP domain and an EL-domain.
By Theorem 15.27a), R[t] is an ACCP domain. Now, seeking a contradiction, we
suppose that R[t] is not an EL-domain. Among the set of all elements in R[t]
admitting inequivalent irreducible factorizations, let p be one of minimal degree.
We may assume
p = f1 · · · fr = g1 · · · gs ,
where for all i, j, (fi ) 6= (gj ) and
m = deg f1 ≥ deg f2 ≥ . . . ≥ deg fr ,
n = deg g1 ≥ deg g2 ≥ . . . ≥ deg gs ,
7. APPLICATION: THE SCHÖNEMANN-EISENSTEIN CRITERION 285

with n ≥ m > 0. Suppose the leading coefficient of f1 (resp. g1 ) is a (resp. b). Put
q = ap−bf1 xn−m g2 · · · gs = f1 (af2 · · · fr −bxn−m g2 · · · gs ) = (ag1 −bf1 xn−m )g2 · · · gs .
Thus q = 0 implies ag1 = bf1 xn−m . If, however, q 6= 0, then
deg(ag1 − bf1 xn−m ) < deg g1 ,
hence deg q < deg p and q has a unique factorization into irreducibles, certainly
including g2 , · · · , gs and f1 . But then f1 must be a factor of ag1 − bf1 xn−m and
thus also of ag1 . Either way ag1 = f1 h for some h ∈ R[t]. Because a is constant, it
can be factored into a product of prime elements a = p1 · · · pr of R, each of which
remains prime in R[t]: R[t]/(pi ) ∼ = R/(pi )[t] is a domain. Since each pi is constant
and f1 is irreducible, we have pi - f1 for all i and it follows that h = ah2 . So
ag1 = f1 ah2 , or g1 = f1 h2 , contradiction. 
Corollary 15.28. Let R be a UFD and let {ti }i∈I be any set of indetermi-
nates. Then S = R[{ti }] is a UFD.
Proof. S When I is finite, apply Theorem 15.26 and induction. When I is
infinite, S = J R[{tj }], as J ranges over all finite subsets of I: any given polyno-
mial can only involve finitely many indeterminates. For f ∈ S, let J be such that
f ∈ R[{tj }], so f = p1 · · · pr is a factorization into prime elements of R[{tj }]. Any
two factorizations in S would themselves lie in some subalgebra involving finitely
many determinates, so the factorization must be unique. 
So a polynomial ring in infinitely many indeterminates over a field k is a non-
Noetherian UFD. This is an important example to keep in mind: the UFD con-
dition is in many ways a very delicate one, but it can still be satisfied by very “large”

We mention without proof two negative results.


Theorem 15.29.
a) (Roitman [Ro93]) There exists an integrally closed atomic domain R such that
R[t] is not atomic.
b) (Anderson-Quintero-Zafrullah) There exists an EL domain R such that R[t] is
not an EL domain.

7. Application: the Schönemann-Eisenstein Criterion


The most famous criterion for irreducibility of univariate polynomials is named
after Ferdinand Eisenstein [Ei50]. However, the version for polynomials over Z
was proven several years earlier by Theodor Schönemann [Sc45], [Sc46]. For many
years now few anglophone texts have associates Schönemann’s name with this result,
and his contribution might have been in real danger of being forgotten were it not
for the beautiful recent article of Cox [Co11] on the early history of this result.
Nowadays it is common to state and prove a version of Eisenstein’s criterion
with respect to a prime ideal in a UFD. We give a slight generalization:
Theorem 15.30. (Schönemann-Eisenstein Criterion) Let R be a domain with
fraction field K, and let f (t) = ad td + . . . + a1 t + a0 ∈ R[t]. Suppose that there
exists a prime ideal p of R such that ad 6∈ p, ai ∈ p for all 0 ≤ i < d and a0 6∈ p2 .
a) If f is primitive, then f is irreducible over R[t].
b) If R is a GCD-domain, then f is irreducible over K[t].
286 15. FACTORIZATION

Proof. a) Suppose to the contrary that f is primitive and reducible over R[t]:
i.e., there exists a factorization f = gh with g(t) = bm tm + . . . + b1 t + b0 , h(t) =
cn tm +. . .+c1 t+c0 , deg(g), deg(h) < deg(f ) and bm cn 6= 0. Since a0 = b0 c0 ∈ p\p2 ,
it follows that exactly one of b0 , c0 lies in p: say it is c0 and not b0 . Moreover, since
ad = bm cn 6∈ p, cn 6∈ p. Let k be the least index such that ck 6∈ p, so 0 < k ≤ n.
Then b0 ck = ak − (b1 ck−1 + . . . + bk c0 ) ∈ p. Since p is prime, it follows that at least
one of b0 , ck lies in p, a contradiction.
b) Suppose R is a GCD-domain and (seeking a contradiction) that f is reducible
over K[t]. By Corollary 15.25b), we may write f = gh with g, h ∈ R[t] and
deg(g), deg(h) < deg(f ). Then the proof of part a) goes gives a contradiction. 

Corollary 15.31. Let R be a GCD-domain containing a prime element π


(e.g. a UFD that is not a field). Then the fraction field K of R is not separably
closed.
Proof. The element π is prime iff the principal ideal p = (π) is a nonzero
prime ideal. Then π 6∈ p2 , so for all n > 1, Pn (t) = tn − π is Eisenstein with respect
to p and hence irreducible in K[t]. Choosing n to be prime to the characteristic of
K yields a degree n separable field extension Ln := K[t]/(Pn ). 

8. Application: Determination of Spec R[t] for a PID R


Let R be a PID. We wish to determine all prime ideals of the ring R[t]. Let us
begin with some general structural considerations. First, R is a one-dimensional
Noetherian UFD; so by Theorems 8.38, 15.27 and 8.50, R[t] is a two-dimensional
Noetherian UFD. Being a UFD, its height one ideals are all principal. Since it
has dimension two, every nonprincipal prime ideal is maximal. Therefore it comes
down to finding all the maximal ideals.

However, to avoid assuming the Dimension Theorem, we begin with milder hy-
potheses on a prime ideal P, following [R, pp. 22-23].

Namely, let P be a nonzero prime ideal of R[t]. We assume – only! – that P


is not principal. By Theorem 15.1, we are entitled to a prime element f1 of P.
Since P =
6 (f1 ), let f2 ∈ P \ (f1 ). Then gcd(f1 , f2 ) = 1: since gcd(f1 , f2 ) | f1 , the
only other possibility is (gcd(f1 , f2 )) = (f1 ), so f1 | f2 and f2 ∈ (f1 ), contradiction.

first claim Let K be the fraction field of R. The elements f1 and f2 are
also relatively prime in the GCD-domain K[t]. Indeed, suppose that f1 = hg1 ,
f2 = hg2 with h, g1 , g2 ∈ K[t] and h a nonunit. By Gauss’ Lemma, we may write
h = ah0 , g1 = b1 γ1 , g2 = b2 γ2 with a1 , b1 , b2 ∈ K and h0 , γ1 , γ2 primitive el-
ements of R[t]. Again by Gauss’ Lemma, h0 γ1 and h0 γ2 are also primitive, so
f1 = hg1 = (ab1 )(h0 γ1 ) ∈ R[t], which implies that ab1 ∈ R. Similarly, ab2 ∈ R, so
h0 is a nonunit of R[t] which divides both f1 and f2 , contradiction.
Let M := hf1 , f2 i, and put m = M ∩ R. It remains to show that, as the
notation suggests, M is a maximal ideal of R[t] and m is a maximal ideal of R.

second claim m 6= 0. Since K[t] is a PID and f1 , f2 are relatively prime in


K[t], there exist a, b ∈ K[t] such that af1 + bf2 = 1. Let 0 6= c ∈ R be an el-
ement which is divisible by the denominator of each coefficient of a and b: then
9. POWER SERIES RINGS OVER UFDS 287

(ca)f1 + (cb)f2 = c with ca, cb ∈ R, so that c ∈ m.

Now put p = P ∩ R, so p = (p) is a prime ideal of the PID R. Moreover,


p = P ∩ R ⊃ M ∩ R = m ) 0,
so p is maximal. Since P ⊃ p, P corresponds to a prime ideal in R[t]/pR[t] =
(R/p)[t], a PID. Therefore P is generated by p ∈ p and an element f ∈ R[t] whose
image in (R/p)[t] is irreducible.

We therefore have proved:


Theorem 15.32. Let R be a PID, and P ∈ Spec R[t]. Then exactly one of the
following holds:
(0) P has height 0: P = (0).
(i) P has height one: P = (f ), for a prime element f ∈ R[t].
(ii) P has height two: P = hp, f i, where p is a prime element of R and f ∈ R[t] is
an element whose image in (R/p)[t] is irreducible. Moreover both P and p := P ∩ R
are maximal, and [R[t]/P : R/p] < ∞.
Exercise 15.13. a) Suppose R has only finitely prime ideals, so is not a Hilbert-
Jacobson ring. By Theorem 12.21, there m ∈ MaxSpec R[t] such that m ∩ R = (0).
Find one, and explain where m fits in to the classification of Theorem 15.32.
b) (Zanello [Za04]) Deduce: for a PID R, the following are equivalent:
(i) R has infinitely many prime ideals.
(ii) Every maximal ideal of R[t] has height two.

9. Power series rings over UFDs


Exercise 15.14. Show: if R is ACCP, so is R[[t]].
In particular, if R is a UFD, R[[t]] is ACCP. But of course the more interesting
question is the following: Must R[[t]] be a UFD?
In contrast to Gauss’s Theorem, whether a formal power series ring over a UFD
must be a UFD was a perplexing problem to 20th century algebraists and remained
open for many years. Some special cases were known relatively early on.
Theorem 15.33. If R is a PID, then R[[t]] is a UFD.
Proof. By Theorem 15.1, it suffices to show that every nonzero prime ideal
P of R[[t]] has a prime element. If t ∈ P, we’re done. Otherwise, let q : R[[t]] → R
be the quotient map and p = q∗ P. Since R is a PID, p can be generated by one
element, and then by Theorem 8.40a), so can P. 
In particular, if k is a field then k[[t1 ]] is a PID, so k[[t1 , t2 ]] = k[[t1 ]][[t2 ]] is a UFD.
This gives the first two cases of the following result, which we will not prove here.
Theorem 15.34. (Rückert [Rü33], Krull [Kr37]) Let k be a field, and let n
be a positive integer. Then k[[t1 , . . . , tn ]] is a UFD.
A significant generalization was proved by Buchsbaum and Samuel, independently,
in 1961. A Noetherian domain R is regular if for every maximal ideal m of R, the
height of m is equal to the dimension of m/m2 as a vector space over the field R/m.
Theorem 15.35. ([Bu61], [Sa61]) If R is a regular UFD, then so is R[[t]].
288 15. FACTORIZATION

The paper [Sa61] also exhibits a UFD R for which R[[t]] is not a UFD.

What about formal power series in countably infinitely many indeterminates? Let
k be a field. There is more than one reasonable way to define such a domain. One
the one hand, one could simply take the “union” (formally, direct limit) of finite
formal power series rings k[[t1 , . . . , tn ]] under the evident inclusion maps. In any
element of this ring, only finitely many indeterminates appear. However, it is useful
also to consider a larger ring, in which the elements are infinite formal k-linear com-
binations of monomials ti1 · · · tin . Let us call this latter domain k[[t1 , . . . , tn , . . .]].
In fact, we have seen this domain before: it is isomorphic to the Dirichlet ring
Dk of functions f : Z+ → k with pointwise addition and convolution product. To
see this, we use unique factorization Q in Z! Namely, enumerate the prime numbers
∞ ai
{pi }∞
i=1 and write n ∈ Z+
as n = i=1 pi , where ai ∈ Z
+
and ai = 0 for all
sufficiently
P large i.
Q∞ ai Then the map which sends f ∈ Dk to the formal power series
n∈Z + f (n) i=1 it gives an isomorphism from Dk to k[[t1 , . . . , t n , . . .]]. In 1959,
E.D. Cashwell and C.J. Everett used Theorem 15.34 to prove the following result.
A key part of their proof was later simplified by C.F. Martin, who pointed out the
applicability of König’s Infinity Lemma.

Theorem 15.36. ([CE59], [Ma71])


a) For any field k, the ring of formal power series k[[t1 , . . . , tn , . . .]] is a UFD.
b) In particular, the ring DC = {f : Z+ → C} of arithmetic functions is a UFD.

In almost any first number theory course one studies unique factorization and also
arithmetic functions, including the Dirichlet ring structure (which e.g. leads to
an immediate proof of the Möbius Inversion Formula). That arithmetic functions
are themselves an example of unique factorization is however a very striking result
that does not seem to be well-known to most students or practitioners of number
theory. I must confess, however, that I (a number theorist) know of no particular
application of Theorem 15.36. I would be interested to learn of one!

10. Nagata’s Criterion


Proposition 15.37. Let R be a domain, S a saturated multiplicative subset,
and f ∈ R \ S. If f is prime as an element of R, it is also prime as an element of
RS .

Proof. Since f ∈ R \ S, f is not a unit in RS . Let α, β ∈ RS be such that


f | αβ in RS . So there exists γ ∈ RS such that γf = αβ; putting α = xs11 , β = xs22 ,
γ = xs33 and clearing denominators, we get s1 s2 x3 f = s3 x1 x2 , so f | r3 x1 x2 . If
f | s3 , then since S is saturated, f ∈ S, contradiction. So, being prime, f divides
x1 or x2 in R, hence a fortiori in RS and therefore it also divides either xs11 or xs22
in RS , since these are associates to x1 and x2 . 

Theorem 15.38. Every localization of a UFD is again a UFD.

Exercise 15.15. Prove Theorem 15.38.


(Suggestions: one gets an easy proof by combining Theorem 15.1 with Proposition
15.37. But the result is also rather straightforward to prove directly.)
10. NAGATA’S CRITERION 289

A saturated multiplicative subset S of R is primal2 if it is generated by the units


of R and by the prime elements of S.
Lemma 15.39. An irreducible element of a primal subset is prime.
Proof. Suppose S is primal and f ∈ S is irreducible. By definition, there
exists a unit u and prime elements π1 , . . . , πn such that f = uπ1 · · · πn . Since uπ1
is also prime, we may as well assume that u = 1. Then, since f is irreducible, we
must have n = 1 and f = π1 . 
Theorem 15.40. For an atomic domain R, the following are equivalent:
(i) Every saturated multiplicative subset of R is primal.
(ii) R is a UFD.
Proof. Since the set R× of units is trivially generated by the empty set of
prime elements, both conditions hold if R is a field, so let us now assume otherwise.
Assume (i). Then, since R is a factorization domain which is not a field, there
exists an irreducible element f of R. Let S be the saturated multiplicative subset
generated by S, which consists of all units of R together with all divisors of positive
powers f n of f . Since S is primal and strictly contains R× , there must exist a
prime element π which divides f n for some n. In other words, f n ∈ πR, and since
πR is prime, we must have that f = xπ for some x ∈ R. Since f is irreducible we
must have x ∈ R× , i.e., f ∼ π and is therefore a prime element. So R is an ACCP
domain and an EL-domain and hence a factorization domain by Theorem 3.3.
Assume (ii), let S be a saturated multiplicative subset of R, and suppose that
f ∈ S \ R× . Then f = uπ1a1 · · · πnan where the πi ’s are prime elements. Since each
πi | f , πi ∈ S for all i. It follows that indeed S is generated by its prime elements
together with the units of R. 
Because of Theorem 15.40, it is no loss of generality to restate Theorem 15.38 as:
the localization of a UFD at a primal subset is again a UFD. The following elegant
result of Nagata may be viewed as a converse.
Theorem 15.41. (Nagata [Na57]) Let R be a factorization domain and S ⊂ R
a primal subset. If the localized domain RS is a UFD, then so is R.
Proof. By Theorem 15.8 it’s enough to show: if f ∈ R is irreducible, then f
is prime.
Case 1: f 6∈ S, so f is not a unit in RS . Since RS is a UFD, it is enough to show
that f is irreducible in RS . So assume not: f = xs11 · xs22 with x1 , x2 ∈ R \ S and
s1 , s2 ∈ S. hen s1 s2 f = x1 x2 . By assumption, we may write s1 = up1 · · · pm and
s2 = vq1 · · · qn , where u, v ∈ R× and pi , qj are all prime elements of R. So p1 | x1 x2 ;
since p1 is a prime, we must have either xp11 ∈ R or xq22 ∈ R. Similarly for all the
other pi ’s and qj ’s, so that we can at each stage divide either the first or the second
factor on the right hand side by each prime element on the left hand side, without
1 x1 x2
leaving the ring R. Therefore we may write f = ( uv ) t1 t2 where t1 , t2 are each
products of the primes pi and qj , hence elements of S, and also such that t1 | x1 ,
t2 | x2 , i.e., the factorization takes place in R. Moreover, since xi ∈ R \ S and
ti ∈ S, xtii is not even a unit in RS , hence a fortiori not a unit in R. Therefore we
have exhibited a nontrivial factorization of f in R, contradiction.
Case 2: f ∈ S. Since S is primal, by Lemma 15.39, f is prime. 
2This terminology is my invention: do you like it?
290 15. FACTORIZATION

Remark: If S is the saturation of a finitely generated multiplicative set, the hy-


pothesis that R is a factorization domain can be omitted.

Application: Let A be a UFD and consider R = A[t]. Put S = A \ {0}. As


for any multiplicative subset of a UFD, S is generated by prime elements. But
moreover, since A[t]/(πA[t]) ∼ = (A/πA)[t], every prime element π of A remains
prime in A[t], so viewing S as the multiplicative subset of A[t] consisting of nonzero
constant polynomials, it too is generated by prime elements. But if F is the fraction
field of A, RS = (A[t])S = F [t] which is a PID and hence a UFD. Nagata’s the-
orem applied to R and S now tells us – for the third time! – that R = A[t] is a UFD.

Nagata used Theorem 15.41 to study the coordinate rings of affine quadric cones.
Let k be a field of characteristic different from 2, and let f (x) = f (x1 , . . . , xn ) ∈
k[x1 , . . . , xn ] be a quadratic form, i.e., a homogeneous polynomial of degree
2 with k coefficients. We assume that f the associated bilinear form (x, y) 7→
1
2 (f (x + y) − f (x) − f (y)) is nonsingular. Equivalently, by making an invertible lin-
ear change of variables every quadratic form can be diagonalized, and a quadratic
form is nonsingular iff it admits a diagonalization
(35) f (x) = a1 x21 + . . . + an x2n with a1 , . . . , an ∈ k × .
We wish to study the affine quadric cone associated to f , namely Rf = k[x]/(f ).
If quadratic forms f and g are isometric – i.e., differ by an invertible linear change
of variables – then Rf ∼= Rg , so we assume if we like that f is in diagonal form
as in (35) above. If n ≥ 3 then every nonsingular diagonal quadratic polynomial
is irreducible, so Rf is a domain. If k is quadratically closed – i.e., admits no
proper quadatic extension – then conversely any binary (n = 2) quadratic form
is reducible, so Rf is not a domain. (If f is not quadratically closed, there exist
irreducible binary quadratic forms, but we will not consider them here.)
Theorem 15.42. Let f = f (x1 , . . . , xn ) ∈ C[x1 , . . . , xn ] be a nondegenerate
quadratic form. Then Rf = k[x]/(f ) is a UFD iff n ≥ 5.
Proof. By the remarks above, Rf is a domain iff n ≥ 3, so we may certainly
restrict to this case. Because C is algebraically closed, every quadratic form in
n ≥ 2 variables is isotropic, i.e., there exists 0 6= a ∈ k n such that f (a) = 0:
indeed, the first n − 1 coordinates of a may be chosen arbitrarily. By an elementary
theorem in the algebraic theory of quadratic forms [La06, Thm. I.3.4], we may
make a change of variables to bring f into the form:
f (x) = x1 x2 + g(x3 , . . . , xn ).
Case 1: Suppose n = 3, so that
f (x) = x1 x2 − ax23
for some a ∈ k × . In this case, to show that Rf is not a UFD it suffices to
show that the images x1 , x2 , x3 of x1 , x2 , x3 in Rf are nonassociate irreducibles,
for then x1 x2 = ax3 2 exhibits a non-unique factorization! To establish this, regard
k[x1 , x2 , x3 ] as a graded C-algebra in the usual way – with x1 , x2 , x3 each of degree
1 – so that the quotient Rf by the homogeneous ideal (f ) inherits a grading. Since
x1 has degree 1, if it were reducible, it would factor as the product of a degree one
10. NAGATA’S CRITERION 291

element c1 x1 + c2 x2 + x3 x3 + (f ) and a degree zero element r + (f ), and thus


(rc1 − 1)x1 + rc2 x2 + rc3 x3 ∈ (f ).
But the left hand side has degree 1, whereas all nonzero elements in (f ) have degree
2 or higher, so r ∈ C[x]× and therefore the factorization is trivial. The irreducibility
of x2 and x3 is proved in the same way. If x1 ∼ x3 in Rf , then we may divide both
sides of x1 x2 − ax3 2 by x1 and deduce that also x2 ∼ x3 . But in the quotient ring
Rf /(x3 ), x3 maps to 0 and x1 and x2 do not, contradiction. So Rf is not a UFD.
Case 2: Suppose n = 4, so f (x) = x1 x2 + g(x3 , x4 ), where g(x3 , x4 ) is a nonsingular
binary form. Here for the first time we use the full strength of the quadratic closure
of k: since k × = k ×2 , any two nonsingular quadratic forms in the same number of
variables are isometric, so we may assume WLOG that
f (x) = x1 x2 − x3 x4 .
Now we argue exactly as in Case 1 above: in Rf , the images x1 , x2 , x3 , x4 are all
non-associate irredcuble elements, so x1 x2 = x3 x4 is a non-unique factorization.
Case 3: n ≥ 5. Then n − 2 ≥ 3, so g is irreducible in the UFD C[x3 , . . . , xn ], hence
also in C[x2 , x3 , . . . , xn ]. Therefore Rf /(x1 ) = C[x1 , . . . , xn ]/(x1 , f ) = C[x2 , . . . , xn ]/(g)
is a domain, i.e., x1 is a prime element. Moreover,
R[x1 −1 ] = C[x1 , . . . , xn , x−1
1 ]/(x1 x2 − g)

∼ g ∼
= C[x1 , . . . , xn , x−1
1 ]/(x2 − ) = C[x1 , x3 , . . . , xn , x−1
1 ]
x1
is a localization of the UFD C[x1 , x3 , . . . , xn ] hence a UFD. By Nagata’s Criterion
(Theorem 15.41), Rf itself is a UFD. 

Now let k be a field of characteristic not 2 and f ∈ k[x1 , . . . , xn ] a nondegenerate


quadratic form. Without changing the isomorphism class of Rq we may diagonalize
f ; moreover without changing the ideal (f ) we may scale by any element of k × , so
without loss of generality we need only consider forms x21 + a2 x22 + . . . + an x2n .
Theorem 15.43. Let k be a field of characteristic different from 2, and let
f = x21 + a2 x22 + . . . + an x2n
be a nonsingular quadratic form over k. Put Rf = k[x]/(f ).
a) If n ≤ 2 then Rf is not an integrally closed domain.
b) If n = 3, then Rf is a UFD iff f is anistropic: ∀a ∈ k n , f (a) = 0 =⇒
a = 0.
c) Suppose f = x21 − ax22 − bx23 − cx24 .
(i) If a is a square in k, then Rf is a UFD iff −bc is not a square in k.
(ii) If none of a, b, c, −ab, −ac, −bc is a square in k, then Rf is a UFD
iff −abc is not a square.
d) If n ≥ 5, then Rf is a UFD.
Proof. a) If n ≤ 2, Rf is never an integrally closed domain.
b) The proof of Theorem 15.42 goes through to show that if f is isotropic (i.e., not
anisotropic), Rf is not a UFD. The anisotropic case is due to Samuel [Sa64].
Part c) is due to T. Ogoma [O74].
Part d) goes back at least to van der Waerden [vdW39]. In [Na57], M. Nagata
gives a short proof using Theorem 15.41. 
292 15. FACTORIZATION

It is also interesting to consider affine rings of inhomogeneous quadric hypersurfaces.


For instance, we state without proof the following result.
Theorem 15.44. For n ≥ 1, let Rn := R[t1 , . . . , tn+1 ]/(t21 + . . . + t2n+1 − 1) be
the ring of polynomial functions on the n-sphere S n .
a) (Bouvier [Bo78]) If n ≥ 2, then Rn is a UFD.
b) (Trotter [Tr88]) R1 is isomorphic to the ring R[cos θ, sin θ] of real trigonometric
polynomials, in which (sin θ)(sin θ) = (1+cos θ)(1−cos θ)) is an explicit non-unique
factorization into irreducible elements. Hence R1 is not a UFD.

11. The Euclidean Criterion


In this section we give a commutative algebraic generalization of Euclid’s proof that
there are infinitely many prime numbers, following [Cl17a]. Euclid’s result on the
face of it pertains to irreducible elements. It happens that in Z every irreducible
element is prime, but that is a different (and deeper) result of Euclid.

It will be helpful to slightly adjust our terminology: here an atom will be the
principal ideal generated by an irreducible element, so two irreducible elements de-
termine the same atom iff they are associate. It seems more interesting to count
atoms than irreducible elements, since in particular whenever an atomic domain
that is not a field has infinite unit group it necessarily has infinitely many irre-
ducible elements but not necessarily infinitely many atoms.

In fact our generalization works in some domains in which not all nonzero nonunits
factor into irreducibles. A Furstenberg domain is a domain R in which every
nonzero nonunit has an irreducible divisor.

We call a ring semiprimitive if it has zero Jacobson radical: i.e., if the only
element lying in every maximal ideal is zero.
Exercise 15.16. a) Show: every atomic domain is a Furstenberg do-
main.
b) Show: the ring Hol(C) of holomorphic functions on the complex plane is
a semiprimitive Furstenberg domain that is not an atomic domain.
Exercise 15.17. a) Show that for a ring R the following conditions are
equivalent:
(i) There is an infinite sequence {In }∞
n=1 of pairwise comaximal, proper
ideals of R.
(ii) The set MaxSpec R is infinite.
b) Let R be a semiprimitive domain that is not a field. Show: MaxSpec R is
infinite.
Theorem 15.45. (Euclidean Criterion [Cl17a, Thm. 2.3]) Let R be a semiprim-
itive domain that is not a field.
a) There is a sequence {an }∞
n=1 of pairwise comaximal nonunits.
b) Suppose moreover that R is a Furstenberg domain. Then there is a se-
quence {pn }∞
n=1 of pairwise comaximal irreducible elements. In particular
R has infinitely many atoms.
11. THE EUCLIDEAN CRITERION 293

Proof. a) We go by induction on n. Since R is not a field there is a1 ∈ R• \R× .


Having chosen a1 , . . . , an that are pairwise comaximal nonunits, because J(R) = (0)
and R is a domain, there is y ∈ R such that
an+1 := ya1 · · · an + 1 ∈ R• \ R× .
(Because a1 ∈ / R× we cannot have an+1 = 0.) Evidently we have hai , an+1 i = R
for all 1 ≤ i ≤ n.
b) Again we go by induction on n. Since R is a Furstenberg domain and not a field
there is some irreducible element p1 . Having chosen pairwise comaximal irreducible
elements p1 , . . . , pn , because J(R) = 0 there is y ∈ R such that
x := yp1 · · · pn + 1 ∈ R• \ R× .
Because we are in a Furstenberg domain, the element x has an irreducible divisor
pn+1 . Then for all 1 ≤ i ≤ n we have
 
 
x Y
1= pn+1 − y  pj  pi ,
pn+1
j6=i

which shows that hpi , pn+1 i = 1. Comaximal irreducible elements are nonassociate,
so R has infinitely many atoms. 
Corollary 15.46. For any domain R, the domain R[t] has infinitely many
atoms.
Exercise 15.18. Prove Corollary 15.46. (Hint: the most interesting case is
when R is finite!)
Corollary 15.47. Let R be a Furstenberg domain, not a field, such that #R >
#R× . Show: R has infinitely many atoms.
Exercise 15.19. a) Prove Corollary 15.47. (Suggestion: make use of
the fact that #I = #R for every √nonzero ideal of R.)
b) Deduce that if R = Z or if R = Z[ −1] then R has infinitely many atoms.
Exercise 15.20. Let K be a number field – i.e., a finite degree extension of
Q – and let ZK be the integral closure of Z in K. Can you use Theorem 15.45 to
show that ZK has infinitely many atoms?3
Exercise 15.21. Let Z be the ring of all algebraic integers – i.e., the integral
closure of Z in C. Show: Z is a semiprimitive domain that is not a field that
has no irreducible elements. Thus in Theorem 15.45b) the hypothesis that R is a
Furstenberg domain cannot be omitted.
Exercise 15.22. Let R be an integrally closed, Noetherian domain of Krull
dimension one. (In other words, R is a Dedekind domain that is not a field. See
Chapter 20 for more on Dedekind domains. Results from this chapter may be helpful
in solving this exercise.) Show that the following are equivalent:
(i) R is semiprimitive.
(ii) R has infinitely many maximal ideals.
(iii) R has inifnitely many atoms.
3In fact Z has infinitely many principal prime ideals, but this is a rather deep theorem in
K
algebraic number theory.
294 15. FACTORIZATION

Proposition 15.48. Let R be a Noetherian domain of Krull dimension one.


Then R is semiprimitive iff MaxSpec R is infinite. When these conditions hold, R
has infinitely many atoms.
Proof. For any domain R, if MaxSpec R = {m1 , . . . , mn } is finite, then
n
\ n
Y
J(R) = mi ⊃ mi ) (0),
i=1 i=1

so R is not semiprimitive. Conversely, if J(R) ) (0), then maximal ideals of R are


minimal prime ideals of the Noetherian ring R/J(R) hence are finite in nubmer
by Theorem 10.13. Since Noetherian domains are atomic and atomic domains are
Furstenberg, when MaxSpec R is infinite the Euclidean Criterion applies to show
that R has infinitely many atoms. 
Lemma 15.49. In an atomic domain, every prime ideal is generated by irre-
ducible elements.
Proof. Let p be a prime ideal of the atomic domain R, and let {xs }s∈S be a
set of nonzero generators for p. Since R is atomic, we may factor xs = fs,1 · · · fs,ns
into irreducibles. Since fs,1 · · · fs,ns lies in the prime ideal p, some fs,js does. By
replacing {xs }s∈S by {fs,js }s∈S we get a set of irreducible generators for p. 
Theorem 15.50. Let R be an atomic domain with finitely atoms. Then R is
semilocal Noetherian and dim R ≤ 1.
Proof. By Lemma 15.49, every prime ideal of R is generated by irreducibles.
Since replacing a generator by an associate element does not change the ideal gen-
erated, we conclude that every prime ideal is finitely generated – so R is Noetherian
by Cohen’s Theorem (Theorem 4.30) – and that there are only finitely many prime
ideals. If the Noetherian domain R had a nonzero, nonmaximal prime ideal p,
then let m be a maximal ideal strictly containing p. By Corollary 8.48, there are
infinitely many prime ideals q with (0) ( q ( m, contradiction. 
Here is what we have shown so far: if R is an atomic domain, not a field, with
finitely many atoms, then R must be semilocal, Noetherian and of Krull dimension
1. Conversely, if R is semilocal, Noetherian of Krull dimension 1 and integrally
closed, then R has finitely many atoms. It’s natural to attempt to remove the
integral closure hypothesis and thereby get a characterization of atomic domains
with finitely many atoms. However, things are not so simple:
Example 15.51. Let k be a field, and consider the subring
R = k[[t2 , t3 ]] = k + t2 k[[t]]
P∞
of the formal power series ring k[[t]]. For 0 6= f = n=0 an tn ∈ k[[t]], we define
v(f ) to be the least n such that an 6= 0. Then v is a discrete valuation on k[[t]],
and the only nonzero prime ideal of k[[t]] is (t) = {f ∈ R | v(f ) > 0} ∪ {0}. The
element t is integral over R – it satisfies the monic polynomial x2 − t2 – and from
this it follows that the only nonzero prime ideal of R is
m = (tk[[t]]) ∩ R = t2 k[[t]] = {f ∈ R | v(f ) ≥ 2} ∪ {0} = ht2 , t3 i.
and R× = {a0 + n≥2 an tn | a0 6= 0}. We will give a complete description of
P

the irreducibles of R. First we claim that f ∈ R is irreducible iff v(f ) ∈ {2, 3}.
11. THE EUCLIDEAN CRITERION 295

Indeed a nontrivial factorization f = xy involves v(x), v(y) ≥ 2 hence v(f ) ≥ 4;


conversely, if v(f ) ≥ 4 then f = t2 tf2 is a nontrivial factorization. Since k × ⊂ R× ,
every irreducible is associate to one of the form
X
t2 + an tn , (v(f ) = 2 case)
n≥3
or one of the form X
t3 + an tn , (v(f ) = 3 case).
n≥4
Associate elements have the same valuation, so certainly no irreducible P of the first
type is associate to an irreducible of the second type. We claim that t2 + n≥3 an tn
is associate to t2 + n≥3 bn tn iff a3 = b3 and t3 + n≥3 an tn is associate to
P P

t3 + n≥3 bn tn iff a4 = b4 . This can be done by direct computation:


P

(t2 + a3 t3 + a1 4t4 + a5 t5 + . . .)(1 + u2 t2 + u3 t3 + . . .)


= t2 + a3 t3 + (a4 + u2 )t4 + (a5 + a3 u2 + u3 )t5 + . . . ,
so a3 = b3 and there is a unique choice of u2 , u3 , . . . leading to an = bn for all
n ≥ 4. The v(f ) = 3 case is similar. Thus there are precisely 2#k atoms, and
hence a finite number iff k is finite.
In Chapter 22 we will revisit the study of Cohen-Kaplansky domains, i.e.,
atomic domains with finitely many atoms, once we have developed further structure
theory of one-dimensional Noetherian domains.
CHAPTER 16

Principal rings and Bézout domains

A ring R in which every ideal is principal is called a principal ring. If R is


moreover a domain, it is called a principal ideal domain (PID).

1. Principal ideal domains


Proposition 16.1. Let R be a UFD.
a) If every maximal ideal of R is principal, then R is a PID.
b) The ring R is a PID iff dim R ≤ 1.
Proof. a) Let R be a UFD in which every maximal ideal is principal. By
Theorem 4.29 it is enough to show that every prime ideal in R is principal. Let
p ∈ Spec R be a nonzero prime ideal. By Theorem 15.1, p contains a prime element
p. But also p lies in a maximal ideal m = (`), so
(p) ⊂ p ⊂ m = (`).
Thus ` | p; since p is a prime element we get (p) = (`), so p = m = (`) is principal.
b) The proof of part a) shows that in no domain can one have a proper containment
of nonzero principal prime ideals. From this it follows that a PID has dimension
at most 1. Conversely, Theorem 15.1 implies that every height one prime ideal in
a UFD is prinicpal, so if dim R ≤ 1 then every prime ideal in R is principal, so –
again by Theorem 4.29 – we get that R is a PID. 
Exercise 16.1. We develop an alternate proof of Proposition 16.1a) following
W. Dubuque.
a) Let R be a UFD, and let S ⊂ R• be any subset. Show: gcd(S) exists.
b) Let R be a UFD in which all maximal ideals are principal, let I be a nonzero ideal
in R. Show that we may write I = gcd(I)J for an ideal J which is not contained
in any proper principal ideal, and conclude that I = gcd(I) is principal.
We now follow with some very familiar examples.
Proposition 16.2. The integer ring Z is a principal ideal domain.
Proposition 16.3. For any field k, the univariate polynomial ring k[t] is a
principal ideal domain.
Surely the most reasonable way to prove Propositions 16.2 and 16.3 is by exploiting
the division algorithms that both rings are well known to possess. We will consider
Euclidean rings later in this chapter, so for now we take it as a challenge to bring
the theory we have developed so far to bear to give other, less reasonable, proofs!
For a property P of rings, a ring R is residually P if for all nonzero ideals I of R,
the quotient R/I has property P.
Lemma 16.4. a) The ring Z is residually Artinian.
297
298 16. PRINCIPAL RINGS AND BÉZOUT DOMAINS

b) For any field k, the ring k[t] is residually Artinian.


c) If a ring R is residually Artinian, then dim R ≤ 1.
Proof. a) For any nonzero n ∈ Z, the ring Z/nZ is finite and thus Artinian.
b) For any nonzero n(t) ∈ k[t], the ring k[t]/(n(t)) is a k-vector space of dimension
equal to the degree of n(t). So k[t]/(n(t)) is Artinian as a k-module hence a fortiori
is Artinian as a ring.
c) By contraposition: if p0 ( p1 ⊂ p2 are prime ideals, then the chain p1 ⊂ p2
shows that dim R/p1 ≥ 1, so R/p1 is not Artinian. 
Combining Proposition 16.1 and Lemma 16.4 we find that it suffices to show that
the rings Z and k[t] are PIDs it suffices to show that they are UFDs. Indeed we have
already seen that k[t] is a UFD, a special case of Theorem 15.27. As for showing
that Z is a UFD, here are two proofs the reader may not have seen before:

First proof: Indeed a decomposition n = p1 · · · pr corresponds to a composition


series for the Z-module Z/nZ. Since Z/nZ is finite, it is certainly Noetherian
and Artinian, so composition series exist. Moreover the Jordan-Hölder theorem
implies that any two composition series have the same number of terms – i.e.,
r = s = `(Z/nZ) – and that after a permutation the sequences of isomorphism
classes of composition factors become identical.

Second proof (Lindemann [Li33], Zermelo [Ze34]): We prove both the existence
and uniqueness of the factorization by an inductive argument, specifically by appeal
to the well-ordering of the positive integers under ≤.
Existence: let S be the set of integers n > 1 which do not have at least one
prime factorization. We wish to show that S is empty so, seeking a contradiction,
suppose not. Then by well-ordering S has a least element, say N . If N is prime,
then we have found a prime factorization, so suppose it is not prime: that is, we
may write N = N1 N2 with 1 < N1 , N2 < N . Thus N1 and N2 are too small to lie
in S so each have prime factorizations, say N1 = p1 · · · pr , N2 = q1 · · · qs , and then
N = p1 · · · pr q1 · · · qs gives a prime factorization of N , contradiction!
Uniqueness: we claim that the factorization of a positive integer is unique.
Assume not; then the set of positive integers which have at least two different
standard form factorizations is nonempty, so has a least element, say N , where:
(36) N = p1 · · · pr = q1 · · · qs .
Here the pi ’s and qj ’s are prime numbers, not necessarily distinct from each other.
However, we must have p1 6= qj for any j. Indeed, if we had such an equality, then
we could cancel and, by an inductive argument we have already rehearsed, reduce
to a situation in which the factorization must be unique. In particular p1 6= q1 .
Without loss of generality, assume p1 < q1 . Then, if we subtract p1 q2 · · · qs from
both sides of (36), we get
(37) M := N − p1 q2 · · · qs = p1 (p2 · · · pr − q2 · · · qs ) = (q1 − p1 )(q2 · · · qs ).
By the assumed minimality of N , the prime factorization of M must be unique.
However, (37) gives two different factorizations of M , and we can use these to
get a contradiction. Specifically, M = p1 (p2 · · · pr − q2 · · · qs ) shows that p1 | M .
Therefore, when we factor M = (q1 − p1 )(q2 · · · qs ) into primes, at least one of the
prime factors must be p1 . But q2 , . . . , qj are already primes which are different from
2. STRUCTURE THEORY OF PRINCIPAL RINGS 299

p1 , so the only way we could get a p1 factor is if p1 | (q1 − p1 ). But this implies
p1 | q1 , and since q1 is also prime this implies p1 = q1 . Contradiction!
Exercise 16.2. Let R be a ring, and suppose that the univariate polynomial
ring R[t] is a PID. Show: R is a field.
Proposition 16.5. Let R be a Noetherian local ring with a principal maximal
ideal m = (a). Then every nonzero ideal of R is of the form (ai ) for some i ∈ N.
In particular, R is a principal ring.
Proof. The Krull Intersection Theorem gives i mi = i (ai ) = 0. It follows
T T
that for any nonzero r ∈ R, there exists a largest i ∈ N such that r ∈ (ai ), i.e., there
exists s ∈ R such that r = sai . But if s were not a unit then it would lie in m and
thus r would lie in mi+1 , contradiction. So s is a unit and (r) = mi . Thus to every
nonzero element r of I we attach a non-negative integer ir . Now if I is any nonzero
ideal of R, choose a nonzero element r of I with ir minimal among elements of I.
Then I ⊃ (r) = mi , and the other containment follows by minimality of ir . 
Proposition 16.6. For any field k, the formal power series ring k[[t]] is a
PID.
Exercise 16.3. Give several proofs of Proposition 16.6.
(Suggestions: (i) Use Theorem 8.40a). (ii) Use Theorem 8.40b) and Proposition
16.5.) (iii) Let I be a nonzero ideal in k[[t]] and let f ∈ I be an element whose
“order of vanishing” – i..e., the index of the least nonzero Laurent series coefficient
– is minimal among elements of I. Show: I = hf i.)

2. Structure theory of principal rings


Proposition 16.7. Let R be a principal ring.
a) For any multiplicative subset S, the localization RS is principal.
b) For any ideal I of R, the quotient R/I is principal.
Exercise 16.4. Prove Proposition 16.7.
Proposition 16.8. For rings R1 , . . . , Rn , the following are equivalent:
(i) Each Ri is a principal
Qnring.
(ii) The direct product i=1 Ri is a principal ring.
Exercise 16.5. a) Prove Proposition 16.8.
b) Show: no infinite product of nonzero principal rings is a principal ring.
A principal ring (R, m) is special if it is a local Artinian ring, i.e., if it is local
and the maximal ideal is principal and nilpotent. The complete structure of ideals
in special principal rings can be deduced from Proposition 16.5: if n is the least
positive integer such that mn = 0, then the ideals of R are precisely the powers
mi = (π i ) for 0 ≤ i ≤ n.

We now give a structure theorem for principal rings, due originally to Krull [Kr24],
as a byproduct of a striking result of Kaplansky [Ka49].
Theorem 16.9. a) (Kaplansky) Let R be a Noetherian ring in which
each maximal ideal is principal. Then R is the direct product of a finite
number of PIDs and special principal rings.
300 16. PRINCIPAL RINGS AND BÉZOUT DOMAINS

b) (Krull) A ring is principal iff it is a finite direct product of rings, each of


which is either a PID or a special principal ring.
Proof. It suffices to prove part a), for then part b) follows immediately.
Step 1: Since R is Noetherian of dimension at most 1, it has finitely many minimal
primes p1 , . . . , pr , and all other primes are maximal. The irreducible components
of Spec R are V (p1 ), . . . , V (pr ). We claim that the irreducible components are
pairwise disjoint: if not there is P ∈ MaxSpec R containing two distinct minimal
primes, and then RP is a Noetherian local ring with principal maximal ideal and
more than one minimal prime. In particular its ideals are not totally ordered under
inclusion, which contradicts Proposition 16.5.
In other words, the minimal Tr primes are pairwise comaximal. ByNExercise 10.13
there is N ∈ Z+ such that i=1 pN i = (0). Taking Ri := R/pi , the Chinese
Remainder Theorem gives
Yr
R= Ri .
i=1
The maximal ideals of Ri are precisely the pushforwards under the quotient map
of the maximal ideals of R containing pi , so each Ri is a Noetherian ring in which
each maximal ideal is principal and having a unique minimal prime. So it suffices
to show that a Noetherian ring R having a unique minimal prime in which each
maximal ideal is principal is either a PID or a special principal ring.
Step 2: If dim R = 0 then R is Artinian local with principal maximal ideal, so
Proposition 16.5 implies that R is a special principal ring. So suppose dim R = 1.
If nil R = (0) then we are assuming that all the prime ideals of R are principal, so
R is a PID by Theorem 4.29. If nil R ) (0), let x be a nonzero nilpotent element,
and let P be a maximal ideal containing ann(x). Then the image of x in RP is a
nonzero nilpotent. The ring RP is one-dimensional local with principal maximal
ideal PRP = hai, say, so by Proposition 16.5 every nonzero ideal of RP is of the
form hai i for some i ∈ N. In such a ring the only nonzero prime ideal is hai, so
since RP has dimension 1, the zero ideal must be prime and RP must be a domain,
contradicting the existence of nonzero nilpotent elements. 
Having studied PIDs, let us now turn to the case of special principal rings and
principal Artinian rings: to be sure a ring is principal Artinian iff it is a finite prod-
uct of special principal rings.. The most familiar examples of principal Artinian
rings (resp. of special principal rings) are Z/nZ (resp. Z/pa Z for a prime number
p). Any quotient of a PID by a nonzero ideal is a principal Artinian ring (Exercise
16.9a)) and any quotient of a PID by a prime power ideal is a special principal ring.

A ring R is self-injective if R is an injective R-module. By Exercise 3.49d), a


domain is self-injective iff it is a field.
Exercise
Q 16.6. Let {Ri }i∈I be a nonempty family of nonzero rings, and put
R := i∈I Ri . Show: R is self-injective iff Ri is self-injective for all i ∈ I.
Proposition 16.10. A principal Artinian ring is self-injective.
Proof. Every special principal ring is a finite product of local special principal
rings, so by Exercise 16.6 we reduce to the case of a local special principal ring
(R, (π)). By Proposition 16.5 if N is the least positive integer such that π N = 0
then the ideals of R are precisely (π a ) for 0 ≤ a ≤ N . By Baer’s Criterion it is
2. STRUCTURE THEORY OF PRINCIPAL RINGS 301

enough to extend every R-module homomorphism ϕ : (π a ) → R to all of R. The


R-module (π a ) is cyclic with annihilator (π N −a ), hence (π a ) ∼
=R R/(π N −a ), so each
a
such homomorphism is uniquely specified by mapping π to any element of R killed
by π N −a , i.e., to any element of (π a ). Thus ϕ maps π a to xπ a for some x ∈ R, and
we can extend it to [x] : R → R, i.e., y ∈ R 7→ xy. 
Proposition 16.10 implies that for N > 1 the ring Z/N Z is self-injective. It follows
that if A is an N -torsion commutative group and x ∈ A has order N , then there
is a subgroup B of A such that A = hxi ⊕ B. This is precisely what is needed
in order to write a finite commutative group as a direct sum of cyclic subgroups.
More generally, we can recover the following well known structure theorem.
Theorem 16.11. Every finitely generated module over a PID is a finite direct
sum of cyclic modules.
Proof. Let R be a PID, and let M be a finitely generated R-module. As
usual, we have the short exact sequence
0 → M [tors] → M → M/M [tors] → 0.
The module M/M [tors] is finitely generated and torsionfree, hence by Theorem
3.61 is free, i.e., is isomorphic to a finite direct sum of copies of R itself. Since
R is Noetherian and M is finitely generated, the torsion submodule T := M [tors]
is also finitely generated, hence it has nonzero annihilator I = (a). Thus T is a
finitely generated module over the special principal ring R/(a). We claim that it
has an element x with annihilator (a), so that the R-submodule hxi is isomorphic
to R/(a). Assuming the claim for the moment, we have an R-submodule T2 of T
such that T = hxi ⊕ T 0 . The submodule T 0 , also being a quotient of T , is again a
finitely generated torsion submodule, so we can argue as above splitting off direct
summands. Since hx1 i ( hx1 i ⊕ hx2 i ( an ascending chain of submodules in the
Noetherian module T , it must terminate, showing that T is a finite direct sum of
cyclic modules.
To see that T has an element annihilated by a: we can easily reduce – e.g.
using the CRT for modules; we leave the details to the reader – to the case of
a = π k a prime power, and then the result is clear, since the annihilator of T is
the ideal generated by the annihilators of the elements of T , so if the annihilator
of each element of T were contained in (π k+1 ) then the annihilator of T would be
contained in (π k+1 ), a contradiction. 
Exercise 16.7. Let R be a PID, and let M be an R-module.
a) Show: M has finite length iff it is a finitely generated torsion R-module.
b) Suppose M is finitely generated. Show: M is indecomposable iff M is free
of rank 1 or M is cyclic with prime power annihilator.
c) Let T be a finitely generated torsion R-module. Use the Krull-Schmidt
Theorem to show that T admits a decomposition as a finite direct sum of
cyclic modules with prime power annihilator. Explain the sense in which
this decomposition is unique. Check the fine print of the structure theorem
for finitely generated modules over a PID in a standard algebra text, and
confirm that Theorem 16.11 and this exercise recover that full result.
Earlier we said that quotients of PIDs are the main examples of special principal
rings. In fact they are the only examples:
302 16. PRINCIPAL RINGS AND BÉZOUT DOMAINS

Theorem 16.12. (Hungerford [Hu68]) Every special principal ring is the quo-
tient of a PID.
Hungerford’s Theorem relies on Cohen’s structure theorem for complete Noetherian
local rings, a topic which we unfortunately do not treat in this text, so the proof
must be omitted.
Exercise 16.8. Show: for a commutative ring R, the following are equivalent:
(i) R is a special principal ring.
(ii) R is the quotient of a discrete valuation ring by a nonzero ideal.
Exercise 16.9.
a) Show that a PID is residually Artinian principal: i.e., every quotient of a
PID by a nonzero ideal is an Artinian principal ring.
b) Hungerford’s Theorem says that the quotients of PIDs by prime power
ideals are precisely the local Artinian principal ring. Show: not every
Artinian principal ring is a quotient of a PID.
(Suggestion: use Exercise 8.41.)
Exercise 16.10. Let k be a field. Let R = k[x, y], and m be the maximal ideal
hx, yi in R. Show that the quotient ring S = R/m2 is nonprincipal. In particular,
if k is finite, then S is a finite nonprincipal local ring.
Exercise 16.11. For n ∈ Z+ , show that the following are equivalent:
(i) Every finite ring of order n is principal.
(ii) The integer n is cubefree, i.e., it is not divisible by the cube of any prime
number.
Exercise 16.12. It follows from the previous exercise that the least cardinality
of a nonprincipal ring is 8.
a) Show that up to isomorphism there are 10 rings of order 8, of which two
are nonprincipal: Z/2Z[x, y]/hx2 , xy, y 2 i and Z/4Z[x]/h2x, x2 i.
b) Show that there is up to isomorphism a unique noncommutative ring of
order 8.
c) Show that for any prime p ≥ 3 there are up to isomorphism 11 rings of
order p3 and one noncommutative ring of order p3 . How many of these
rings are principal?
Exercise 16.13. (Inspired by http://math.stackexchange.com/questions/361258)
Show: for a ring R the following are equivalent:
(i) The polynomial ring R[t] is principal.
(ii) R is a finite product of fields.
(Suggestions: (ii) =⇒ (i) is easy. As for (i) =⇒ (ii), show first that if R[t] is
principal, then R is principal Artinian, and then reduce to showing that if R is a
local principal Artinian ring with nonzero maximal ideal p = hπi, then R[t] is not
principal. For this let m := hπ, ti ∈ MaxSpec R[t] and show that dimR[t]/m m/m2 =
dimR/πR m/m2 = 2.)

3. Euclidean functions and Euclidean rings


THIS SECTION WILL SOON BE REWRITTEN
3. EUCLIDEAN FUNCTIONS AND EUCLIDEAN RINGS 303

If R is a domain, then a Euclidean function is a function N : R \ {0} → N


such that: for all a ∈ R and b ∈ R \ {0}, there exists q, r ∈ R such that a = qb + r
with r = 0 or N (r) < N (b).
Exercise 16.14. For each of the following rings R, we give a function N :
R \ {0} → N. Verify that N is a Euclidean function on R.
a) R = Z, N (a) = |a|.
b) R = k[t], N (p(t))
P∞= deg(p(t)).
c) R = k[[t]], N ( n=0 an tn ) := the
√ least n such that an 6= 0. √
d) For d ∈ {−2, −1, 2, 3}, R = Z[ d] = Z[t]/(t2 − d), N (a + b d) = |a2 − db2 |.
Exercise 16.15. Let R be a local PID. Use Proposition 16.5 to construct a
Euclidean function on R.
The virtue of Euclidean functions lies in the following result:
Proposition 16.13. A domain R which admits a Euclidean function is a PID.
Proof. Let I be any nonzero ideal of R, and choose an element x in I such
that N (x) is minimal. For any y ∈ I, by definition of a Euclidean function we may
write y = qx + r with r = 0 or N (y) < N (x). Our choice of x rules out the second
alternative and thus we have y = qx, i.e., y ∈ (x). 
Remark: It is common to define a Euclidean domain as a domain which admits
some Euclidean function. In my opinion, this definition is somewhat of a false step
in the theory. The merit of the notion of a Euclidean function is that for certain
rings R there is an evident Euclidean function (e.g. for Z and k[t]), and when
this occurs one deduces immediately that R is a PID. Moreover, if the Euclidean
function is (in some sense) computable, the Euclidean algorithm can be used to
find greatest common divisors and, accordingly, generators of ideals.
In contrast, since the specific Euclidean function is not part of the definition
of a Euclidean domain, in order to show that a given ring R is not Euclidean one
needs to argue that no Euclidean function can exist. Evidently if R is not a PID,
no such function can exist. However, there are PIDs which admit no Euclidean
function (more details are given below).

One way to assess the naturality of the notion Euclidean domain is to examine
its stability under small perturbations of the definition. In other words, to what
extent do similar axioms for a “Euclidean function” lead to an equivalent class of
rings?

There are several results to the effect that a certain weakening of the definition
of a Euclidean function captures precisely the class of all PIDs. For example:
Theorem 16.14. (Dedekind-Hasse) For a domain R with fraction field K, the
following are equivalent:
(i) There exists a function N : K → Q satisfying:
(QN1) ∀x ∈ K, N (x) ≥ 0; N (x) = 0 ⇐⇒ x = 0.
(QN2) ∀x, y ∈ K, N (xy) = N (x)N (y).
(QN3) ∀x ∈ R, N (x) ∈ Z.
(QN4) ∀x ∈ R, N (x) = 1 ⇐⇒ x ∈ R× .
(QN5) ∀x ∈ K \ R, ∃ a, b ∈ R such that 0 < N (ax − b) < 1.
(ii) R is a principal ideal domain.
304 16. PRINCIPAL RINGS AND BÉZOUT DOMAINS

Exercise 16.16. Prove Theorem 16.14. (Suggestions: (i) =⇒ (ii) is the usual
argument that an ideal is generated by any element of minimal norm. Conversely, if
R is a PID, define N on R \ {0} by, e.g., setting N (x) to be 2r , where r = `(R/xR)
is the number of irreducibles appearing in a factorization of x and extend this map
to K.)
In another direction, P. Samuel considered the notion of a W-Euclidean function
on a domain R. Here W is a well-ordered set and N : R → W is a function such
that for all a ∈ R, b ∈ R \ {0} such that b - a, ∃ q, r ∈ R with a = qb + r and
N (r) < N (b). If R admits, for some W , a W -Euclidean function, then R is a PID.
Exercise 16.17. Show: a domain is W -Euclidean for some finite W iff it is a
field.
Say that a domain is Samuel-Euclidean if it is W -Euclidean for some √ well-ordered
set W . Samuel remarks that the an imaginary quadratic fields Q( −d) has a
Samuel-Euclidean ring of integers Rd iff d = 1, 2, 3, 7, 11. On the other hand, it
goes back at least to Gauss that for each of d = 19, 43, 67, 163 the ring Rd is a PID.
Thus there are PIDs which are not Samuel-Euclidean. Samuel further showed that
any Samuel-Euclidean ring is W -Euclidean for a unique minimal well-ordered set
(up to canonical order isomorphism) WR and asked the question of whether one has
WR ≤ N for all domains R. This was answered in the negative by Hiblot [Hi75],
[Hi77].

4. Bézout domains
Proposition 16.15. Let a, b be elements of a domain R. If the ideal ha, bi is
principal, then its generator is a greatest common divisor of a and b.
Proof. In other words, we are assuming the existence of some d ∈ R such
that dR = aR + bR. Then a, b ∈ dR, so d is a common divisor of a and b. If e | a
and e | b then since there are x, y ∈ R such that d = xa + yb, we have e | d. 
Corollary 16.16. For a domain R, the following are equivalent:
(i) Every finitely generated ideal is principal.
(ii) For any two elements a and b of R, their gcd exists and is an R-linear
combination of a and b.
Proof. (i) =⇒ (ii) is immediate from Proposition 16.15: gcd(a, b) will be a
generator of the ideal ha, bi. Conversely, if d = gcd(a, b) exists and is of the form
d = xa + yb for some x, y ∈ R, then clearly (d) = ha, bi, so that every ideal with
two generators is principal. By an obvious induction argument, we conclude that
any finitely generated ideal is principal. 
At least according to some, it was Étienne Bézout who first explicitly noted that
for polynomials P, Q ∈ k[t], gcd(a, b) exists and is a linear combination of a and
b: this fact is called Bézout’s identity or Bézout’s Lemma. For this (some-
what tenuous) reason, a possibly non-Noetherian domain satisfying the equivalent
conditions of Corollary 16.16 is called a Bézout domain.
Exercise 16.18. Show: a localization of a Bézout domain is again a Bézout
domain.
Theorem 16.17. For a Bézout domain R, the following are equivalent:
4. BÉZOUT DOMAINS 305

(i) R is a PID.
(ii) R is Noetherian.
(iii) R is a UFD.
(iv) R is an ACCP domain.
(v) R is an atomic domain.
Proof. (i) ⇐⇒ (ii) immediately from the definitions.
(i) =⇒ (iii): this is Corollary 15.2.
(iii) =⇒ (iv) =⇒ (v) holds for all domains.
(v) =⇒ (iii): A Bézout domain is a GCD-domain (Corollary 16.16), a GCD-
domain is an EL-domain (Proposition 15.11), and an atomic EL-domain is a UFD
(Theorem 15.8), so a Bézout atomic domain is a UFD.
(iv) =⇒ (ii): assume that R is not Noetherian. Then it admits an infinitely
generated ideal I, which we can use to build an infinite strictly ascending chain of
finitely generated ideals I1 ( I2 ( . . . ( I. Since R is Bézout, each Ii is principal,
contradicting ACCP. 
Let us say that a domain is properly Bézout if it is Bézout but not a PID.

We have already seen some examples of properly Bézout domains: the ring of
entire functions (Theorem 5.21) and the ring of all algebraic integers (Theorem
5.1). To get further examples we move on to the next topic: valuation rings.
CHAPTER 17

Valuation rings

1. Basic theory
Consider the divisibility relation – i.e., a | b – on a domain R. Evidently it is
reflexive and transitive, so is a quasi-ordering.1 Divisibility need not be a par-
tial ordering because a | b and b | a does not imply that a = b but only that a
and b are associates: (a) = (b). However, one of the first ideas of ideal theory is
to view associate elements as being somehow “equivalent.” This motivates us to
consider the equivalence relation on R in which a ∼ b iff (a) = (b). This is easily
seen to be a monoidal equivalence relation. In plainer language, if (a1 ) = (a2 )
and (b1 ) = (b2 ), then (a1 b1 ) = (a2 b2 ). We can therefore consider the commutative
monoid of principal ideals of R under multiplication, on which the divisibility rela-
tion is a partial ordering.
Having made a quasi-ordering into a partial ordering, it is natural to ask for
conditions under which the divisibility relation induces a total ordering. Equiv-
alently, for any a, b ∈ R either a | b or b | a.
Proposition 17.1. Let R be a domain with fraction field K. the following are
equivalent:
(i) For every a, b ∈ R, a | b or b | a.
(ii) For every 0 6= x ∈ K, x ∈ R or x−1 ∈ R.
Exercise 17.1. Prove Proposition 17.1.
A domain R satisfying the conditions of Proposition 17.1 is called a valuation
domain or valuation ring.

Note that any field is a valuation ring. This is a trivial example which is often
implicitly excluded from consideration (we will try our best to be explicit in our
exclusion of trivial cases). Apart from this, in a first algebra course one may not
see examples of valuation rings. But we have: if p is a prime number, then the ring
Z(p) of integers localized at p is such an example. Define x |p y if ordp ( xy ) ≥ 0. Then
p-divisibility is immediately seen to be a total quasi-ordering: given two integers,
at least one p-divides the other. The fundamental theorem of arithmetic implies
x | y ⇐⇒ ∀ primes p, x |p y.
However, in Z(p) , we have x | y ⇐⇒ x |p y, i.e., we have localized the divisibility
relation to get a total quasi-order: Z(p) is a valuation domain.

This argument generalizes as follows: let R be a PID2 and p = (π) be a prime


1By definition, a quasi-ordering is a reflexive, transitive binary relation on a set.
2In fact we can take R to be any Dedekind domain, as soon as we know what such a thing
is. See §18.
307
308 17. VALUATION RINGS

ideal of R. We define ordp (x) to be the least n such that (x) ⊃ pn , and extend it
to a map on K × by ordp ( xy ) = ordp (x) − ordp (y). (One should check that this is
well-defined; this is easy.) Finally, we define x |p y to mean ordp ( xy ) ≥ 0. Arguing
as above, we see that the localization Rp is a valuation ring.

In showing that Rp was a valuation domain we proceeded by constructing a map


ordp on the nonzero elements of the fraction field K. This can be generalized, as
follows: if R is a domain with quotient field K, we can extend the divisibility re-
lation to K × by saying that x | y iff xy ∈ R. Clearly x | y and y | x iff xy is a unit
in R. Therefore the quotient of (K × , ·) on which divisibility (from R!) becomes a
partial ordering is precisely the quotient group K × /R× .

For [x], [y] ∈ K × /R× , let us write [x] ≤ [y] if [ xy ] ∈ R. (Take a second and
check that this is well-defined.)
Exercise 17.2. Show: the divisibility quasi-ordering on R is a total ordering
iff the ordering on K × /R× is a total ordering.
In other words, if R is a valuation ring, then the canonical map v : K × → K × /R×
is a homomorphism onto a totally ordered commutative group. Let us relabel the
quotient group by G and denote the group law by addition, so that the homomor-
phism property gets recorded as

(VRK1) ∀x, y ∈ K × v(ab) = v(a) + v(b).

We recover R as
R = {x ∈ K × | v(x) ≥ 0} ∪ {0}.
Everything that has been said so far takes into account only the multiplicative
structure on R. So the following additional property is very important:

(VRK2) ∀x, y ∈ K × | x + y 6= 0, v(x + y) ≥ min(v(x), v(y)).


y x+y y
Indeed, suppose WLOG that v(x) ≤ v(y), i.e., x ∈ R. Then x = 1+ x ∈ R so
v(x) ≤ v(x + y).
Exercise 17.3. Suppose v(x) 6= v(y). Show that v(x + y) = min(v(x), v(y)).
Let (G, +, ≤) be a totally ordered commutative group. We write G+ = {g ∈ G | g ≥
0}, so G+ is a totally ordered submonoid of G. A (G-valued) valuation on a
field K is a surjective map v : K × → G satisfying (VRK1) and (VRK2) above.
Exercise 17.4. Let v : K × → G be a valuation. Let R be the set of elements
of K × with non-negative valuation, together with 0. Show: R is a valuation ring
with fraction field K.
Exercise 17.5. Let R be a domain, G a totally ordered group and v : R\{0} →
G+ be a map which satisfies all of the following properties:
(VRR1) ∀x, y ∈ R \ {0}, v(ab) = v(a) + v(b).
(VRR2) ∀x, y ∈ R \ {0} | x + y 6= 0, v(x + y) ≥ min(v(x), v(y)).
(VRR3) v(R \ {0}) ⊃ G+ .
Show that there is a unique extension of v to a valuation v : K × → G, namely
v(x/y) = v(x) − v(y).
2. ORDERED COMMUTATIVE GROUPS 309

Proposition 17.2. A valuation ring is a local domain, in which the unique


maximal ideal m consists of elements of positive valuation.
Proof. The set of elements of strictly positive valuation form an ideal m of
R. Its complement, the set of elements of zero valuation, is the group of units. 

Proposition 17.3. Let I be a finitely generated ideal in a valuation ring. Then


the set v(I) has a least element, and for any x ∈ I of minimal valuation, I = v(x).
Proof. Write I = hx1 , . . . , xn i with v(x1 ) ≤ . . . ≤ v(xn ). Then for any
r1 , . . . , rn ∈ R, v(r1 x1 + . . . + rn xn ) ≥ mini v(ri xi ) ≥ v(x1 ), so x1 is an element of
I of minimal valuation. Thus for all i > 1, v( xx1i ) ≥ 0, so xx1i ∈ R and x1 | xi . So
I = hx1 i. 

Corollary 17.4. A valuation ring is a Bézout domain. In particular, a Noe-


therian valuation ring is a PID.
Conversely:
Proposition 17.5. A local Bézout domain is a valuation domain.
Proof. Let x, y be elements of a local Bézout domain, and suppose d =
gcd(x, y). Then xd and yd are coprime. Since in a local ring the nonunits form an
ideal, this implies that at least one of xd and yd is a unit. In other words, up to
associates, d = x (so x | y) or d = y (so y | x). 

A valued field is a field together K together with a valuation v : K × → G. An


isomorphism of valued fields (K, v) → (K 0 , v 0 ) is an isomorphism f : K → K 0 of
fields such that v = v 0 ◦f . A valued field is trivial if G is the trivial group. Evidently
each field carries a unique trivial valuation up to isomorphism: the valuation ring
is K itself. The reader will lose nothing by making the tacit assumption that all
valuations are nontrivial.
Exercise 17.6. A chain ring is a ring R in which the partially ordered set
I(R) of ideals of R is linearly ordered.
a) Show: for a ring R, the following are equivalent:
(i) R is a chain ring.
(ii) For all x, y ∈ R, either xR ⊂ yR or yR ⊂ xR.
b) Show: a domain R is a chain ring iff it is a valuation ring.

2. Ordered commutative groups


Let (G, +) be an commutative group, written additively. In particular the identity
element of G will be denoted by 0. As for rings, we write G• for G \ {0}.

By an ordering on G we mean a total (a.k.a. linear) ordering ≤ on G which


is compatible with the addition law in the following sense:

(OAG) For all x1 , x2 , y1 , y2 ∈ G, x1 ≤ x2 and y1 ≤ y2 implies x1 + y1 ≤ x2 + y2 .

One has the evident notions of a homomorphism of ordered commutative groups,


namely an isotone group homomorphism.
310 17. VALUATION RINGS

Exercise 17.7. Let (G, ≤) be an ordered commutative group.


a) Let x ∈ G• . Show: either x > 0 or −x > 0 but not both.
b) Show: for all x, y ∈ G, x ≤ y ⇐⇒ −y ≤ −x.
Exercise 17.8. Let (G, ≤) be an ordered commutative group, and let H be a
subgroup of G. Show: the induced order on H makes H into an ordered commutative
group.
Example 17.6. For an ordered field (F, ≤), the additive group (F, +) is an
ordered commutative group. In particular, the additive group (R, +) of the real
numbers is an ordered commutative group, as is any subgroup. In particular, (Z, +)
and (Q, +) are ordered commutative groups.
Exercise 17.9. Exhibit an commutative group which admits two nonisomor-
phic orderings.
Example 17.7. (Lexicographic ordering): Let {Gi }i∈I be a nonempty indexed
family of ordered commutative groups. Suppose that we are givenQa well-ordering
on the index set I. We may then endow the direct product G = i∈I Gi with the
structure of an ordered commutative group, as follows: for (gi ), (hi ) ∈ G, we decree
(gi ) < (hi ) if for the least index i such that gi 6= hi , gi < hi .
Q
Exercise 17.10. Check that the lexicographic ordering on the product i∈I Gi
is indeed a total ordering on G.
Theorem 17.8. (Levi [Le43]) For an commutative group G, the following are
equivalent:
(i) G admits at least one ordering.
(ii) G is torsionfree.
Proof. (i) =⇒ (ii) Suppose < is an ordering on G and let x ∈ G• . Then
exactly one of x, −x is positive; without loss of generality say it is x. Then for all
n ∈ Z+ , nx = x + . . . x (n times) must be positive, so x has infinite order in G.
(ii) =⇒ (i): Let G be a torsionfree commutative group. By Corollary 3.95, G is a
flat Z-module. Tensoring the injection Z ,→ Q gives us an injection G ,→L G ⊗ Q.
Since Q is a field, the Q-module G ⊗ Q is free, i.e., it is isomorphic to i∈I Q.
Choose a well-ordering on I. Give each copy ofLQ its standard ordering as a subfield

of R and put the lexicographic ordering on i∈I Q = G ⊗ Q. Via the injection
G ,→ G ⊗ Q this induces an ordering on G. 
Exercise 17.11. a) Show: the commutative group Z admits exactly one order-
ing (here when we say “ordering”, we always mean “ordering compatible with the
group structure in the sense of (OAG).
b) Give an example of an commutative group which admits two distinct – even
nonisomorphic – orderings.
An ordered commutative group (G, +) is Archimedean if for all x, y ∈ G with
x > 0, there exists n ∈ Z+ with nx > y.
Exercise 17.12. a) Suppose H is a subgroup of the Archimedean ordered group
(G, +). Show: the induced ordering on G is Archimedean.
b) Let (G, +) be an ordered commutative group such that there exists an embed-
ding (G, +) ,→ (R, +) into the additive group of the real numbers. Deduce: G is
Archimedean.
2. ORDERED COMMUTATIVE GROUPS 311

Conversely:
Theorem 17.9. (Hölder [Hö01]) Let (G, +) be an ordered commutative group.
If G is Archimedean, there exists an embedding (G, +) ,→ (R, +).
Proof. We may assume G is nontrivial. Fix any positive element x of G. We
will construct an order embedding of G into R mapping x to 1.
Namely, let y ∈ G. Then the set of integers n such that nx ≤ y has a maximal
element n0 . Put y1 = y − n0 x. Now let n1 be the largest integer n such that
nx ≤ 10y1 : observe that 0 ≤ n1 < 10. Continuing in this way we get P a set of integers
∞ nk
n1 , n2 , . . . ∈ {0, . . . , 9}. We define ϕ(y) to be the real number n0 + k=1 10 k . It is
0 0
not hard to show that ϕ is isotone – y ≤ y =⇒ ϕ(y) ≤ ϕ(y ) – and also that ϕ is
injective: we leave these tasks to the reader.
But let us check that ϕ is a homomorphism of groups. For y ∈ G, and r ∈ Z+ ,
let 10nr be the rational number obtained by truncating ϕ(y) at r decimal places.
The numerator n is characterized by nx ≤ 10r y < (n + 1)x. For y 0 ∈ G, if
n0 x ≤ 10r y 0 ≤ (n0 + 1)x, then
(n + n0 )x ≤ 10r (y + y 0 ) < (n + n0 + 2)x,
so
2
ϕ(y + y 0 ) − (n + n0 )10−r <
10r
and thus
4
|ϕ(y + y 0 ) − ϕ(y) − ϕ(y 0 )| <
.
10r
0 0
Since r is arbitrary, we conclude ϕ(y + y ) = ϕ(y) + ϕ(y ). 
A nontrivial ordered commutative group which can be embedded in R is said to
have rank one. For many applications this is by far the most important case.
Later we will give the general definition of the rank of an ordered commutative
group.

The following result gives another characterization of valuation rings of rank one.
Lemma 17.10. Let R be a domain, (G, ≤) a totally ordered commutative group,
and let G+ = {g ∈ G | g ≥ 0}. Then:
a) G+ is an ordered commutative monoid.
b) The monoid ring R[G+ ] is a domain.
c) The group ring R[G] is naturally isomorphic to the localization of R[G+ ] at the
multiplicative subset G+ . In particular, R[G] is a domain.
Exercise 17.13. a) Write out the statements of Lemma 17.10 when G = Z.
b) Prove Lemma 17.10.
Theorem 17.11. (Malcev, Neumann) For any ordered commutative group G,
there exists a valuation domain with value group isomorphic to G.
Proof. It suffices to construct a field K and a surjective map v : K × → G
satisfying (VD1K) and (VD2K). Let k be any field and put A = k[G≥0 ]. By

Lemma 17.10, A is a domain; let P K be its fraction field. Define a map v : A → G
by sending a nonzero element g∈G ag g to the least g for which ag 6= 0. Then v
satisfies the three properties of Exercise 17.5 and therefore extends uniquely to a
valuation v : K × → G, where K is the fraction field of R. 
312 17. VALUATION RINGS

Recall from §5.5 the notion of a “big monoid ring” k[[Γ]], the collection of all
functions f : Γ → k under pointwise addition and convolution product. As we saw
though, in order for the convolution product to be defined “purely algebraically” –
i.e., without recourse to some limiting process – we need to impose the condition of
divisor finiteness on Γ. It follows easily from Proposition 17.20 that for Γ = G≥0 the
monoid of non-negative elements in a totally ordered commutative group, divisor
finiteness holds iff G = Z, i.e., iff the valuation is discrete.
However, Malcev [Ma48] and Neumann [Ne49] independently found a way
around this by considering a set in between k[G≥0 ] and k[[G≥0 ]]. Namely, define
kMN [G≥0 ] to be the set of all functions f : G≥0 → k such that the support of f
– i.e., the set of g ∈ G≥0 such that f (g) 6= 0 – is well-ordered. It turns out that
on such functions the convolution product can be defined and endows kMN [G≥0 ]
with a domain. The fraction field kMN (G) is simply the collection of all functions
f : G → k with well-ordered support. Moreover, mapping each such nonzero
function to the least element of G in its support gives a G-valued valuation. The
elements of such rings are called Malcev-Neumann series.

2.1. Convex Subgroups.

A subset S of a totally ordered set (X, ≤) is convex if for all x < y < z ∈ X,
if x, z ∈ S, then y ∈ S.
Exercise 17.14. Let H be a subgroup of an ordered commutative group (G, ≤).
Show: H is convex iff for all x, y ∈ G with 0 ≤ x ≤ y, if y ∈ H then also x ∈ H.
Proposition 17.12. Let (G, ≤) be an ordered commutative group, and let C(G)
be the family of convex subgroups of G. Then C(G) is totally ordered under inclu-
sion.
Proof. Let H1 and H2 be convex subgroups. Seeking a contradiction, we
suppose there is h1 ∈ H1 \ H2 and h2 ∈ H2 \ H1 . Since subgroups are closed under
inversion, we may assume that h1 , h2 ≥ 0 and then, without loss of generality,
that 0 ≤ h1 ≤ h2 . Since H2 is a convex subgroup, it follows that h1 ∈ H2 ,
contradiction. 

For an ordered commutative group G, we define r(G) to be the order isomorphism


type of the linearly ordered set C(G)• = C(G)\{{0}} of nontrivial convex subgroups
of G. When this set is finite we may view r(G) as a natural number. In particular,
r(G) = 1 iff G is nontrivial and has no proper, nontrivial convex subgroups.
Exercise 17.15. a) Let G1 and G2 be ordered commutative groups, and let
G = G1 × G2 be lexicographically ordered. Show: r(G) = r(G1 ) + r(G2 ), where
on the right hand side we have the ordered sum: every element of the first linearly
ordered set is less than every element of the second linearly ordered set.
b) Let n ∈ Z+ . Show: r(Zn ) = n.
Proposition 17.13. For a nontrivial ordered commutative group G, the fol-
lowing are equivalent:
(i) G is Archimedean.
(ii) r(G) = 1.
Proof. 
3. CONNECTIONS WITH INTEGRAL CLOSURE 313

In view of Proposition 17.13 it makes sense to call r(G) the rank of the linearly
ordered group G: indeed we have already defined a group to have rank one if
it is nontrivial and can be order embedded in (R, +). By Theorem 17.9, G is
Archimedean iff it can be order embedded in (R, +), so by Proposition 17.13 our
new notion of rank coincides with our old notion of rank one.
Theorem 17.14. Let v : K × → (G, ≤) be a valuation on a field K, with
valuation ring R. There is an inclusion reversing bijection Φ : Spec R → C(G)
given by
p 7→ G \ ±v(p).
Proof. Step 1: We claim that for p ∈ Spec R, Φ(p) is a convex subgroup.
Clearly Φ(p) contains 0 and is closed under inversion, so suppose σ1 , σ2 ∈ Φ(p).
Since Φ(p) is closed under inversion, we may assume that either σ1 , σ2 > 0 or
σ1 > 0, σ2 < 0 and σ1 + σ2 > 0.
Case 1: Suppose σ1 , σ2 > 0. Choose x1 , x2 ∈ R with v(xi ) = σi for i = 1, 2. If
σ1 + σ2 ∈/ Φ(p), then there is x ∈ p with v(x) = σ1 + σ2 = v(x1 x2 ). Thus ux = x1 x2
for some u ∈ R× , so x1 x2 ∈ p. Since p is prime this implies that at least one of x1 ,
x2 lies in p, hence at least one of σ1 , σ2 does not lie in Φ(p), contradiction.
Case 2: Suppose σ1 > 0, σ2 < 0, σ1 + σ2 > 0. If σ1 + σ2 ∈ / Φ(p), then there is x ∈ p
with v(x) = σ1 + σ2 . Choose y ∈ R with v(y) = −σ2 . Then yx ∈ p and v(yx) = σ1 ,
so σ1 ∈ v(p), contradiction.
To show convexity: let 0 ≤ σ1 ≤ σ2 ∈ G, and suppose σ2 ∈ Φ(p). If σ1 ∈ v(p),
then there exists x ∈ p with v(x) = σ1 . There is y ∈ R such that v(y) = σ2 − σ1 ,
and thus v(yx) = σ2 , so σ2 ∈ / Φ(p), contradiction.
Step 2: To a convex subgroup H of G, we associate
Ψ(H) = {x ∈ R• | v(x) ∈
/ H} ∪ {0}.
By similar – but easier – reasoning to the above, one checks that Ψ(H) ∈ Spec R.
Step 3: One checks that Φ and Ψ are mutually inverse maps, hence Φ is a bijection.

Exercise 17.16. Supply the details of Steps 2 and 3 in the proof of Theorem
17.14.
Exercise 17.17. Show: the maps Φ and Ψ are obtained by restricting the
Galois connection associated to a relation on R × G.
Corollary 17.15. For a valuation ring R, the following are equivalent:
(i) R has rank one, i.e., the value group is Archimedean.
(ii) R has Krull dimension one.

3. Connections with integral closure


Let (R, mR ) and (T, mT ) be local rings with R ⊂ T . We say that T dominates
R, and write R ≤ T , if mT ∩ R = mR .
Lemma 17.16. Let R be a subring of a field K, and let p ∈ Spec R. Then there
exists a valuation ring T of K such that R ⊂ T and mT ∩ R = p.
Proof. (Matsumura)
Step 0: We may replace R by Rp and thus asume that (R, p) is a local ring. In this
case, what we are trying to show is precisely that there exists a valuation ring of
314 17. VALUATION RINGS

K dominating R.
Step 1: Let F be the set of all rings R0 with R ⊂ R0 ⊂ K such that pR0 ( R0 ,
partially ordered by inclusion. We have R ∈ F, so F 6= ∅. Moreover the union of a
chain in F is again an element of F, so Zorn’s Lemma gives us a maximal element
T of F. Since pT ( T , there exists a maximal ideal m of T containing pT . Since
T ⊂ Tm and Tm ∈ F, by maximality of T we have T = Tm , so (T, m) is a local ring
dominating (Rp , p).
Step 2: We claim that T is a valuation ring.
proof of claim: Let x ∈ K × . We wish to show that at least one of x, x−1 lies in
T . Seeking a contradiction, assume neither is the case. Then T [x] properly contains
T , so by maximality of T we have 1 ∈ pT [x], i.e., we get a relation of the form
1 = a0 + a1 x + . . . + an xn , ai ∈ pT.
Since T is local, 1 − a0 ∈ T × , and the relation may be rewritten in the form
(38) 1 = b1 x + . . . + bn xn , bi ∈ m.
Among all such relations, we may choose one with minimal exponent n. In exactly
the same way, T [x−1 ] properly contains T and thus there exists a relation
(39) 1 = c1 x−1 + . . . + cn x−m , ci ∈ m,
and among all such relations we may choose one with minimal m. Without loss
of generality m ≤ n: otherwise interchange x and x−1 . Then multiplying (39) by
bn xn and subtracting from (38) gives another relation of the form (38) but with
exponent smaller than n, contradiction. 

A subring R of a field K is a maximal subring if R ( K and there is no ring


intermediate between R and K.
Exercise 17.18. a) Show: any field K admits at least one maximal subring.
b) Show: if K is not finite of prime order, then all maximal subrings are nonzero.
c) Excluding the case in which K is finite of prime order, show: any maximal
subring of K is a valuation ring of K (possibly a field).
Exercise 17.19. Let K be a field, and R a valuation ring of K. Show the
following are equivalent:
(i) R is a maximal subring of K.
(ii) R has rank at most one.
Theorem 17.17. Let K be a field and R ⊂ K a subring. Then the integral
closure R of R in K is the intersection of all valuation rings of K containing R.
Proof. Let R be the intersection of all valuation rings of K containing R.
Since each such ring is integrally closed in K and the intersection of a family of
rings each integrally closed in K is again integrally closed in K, R is integrally
closed in K, whence R ⊂ R.
Conversely, let x ∈ K \ R. It suffices to find a valuation ring of K containing R but
not x. Let y = x−1 . The ideal yR[y] of R[y] is proper: for if 1 = a1 y + . . . + an y n
with ai ∈ R, then x would be integral over R. Let p be a maximal ideal of R
containing y. By Lemma 17.16, there exists a valuation ring T of K such that
R[y] ⊂ K and mT ∩ R[y] = p. Then y = x−1 ∈ mT , so x ∈ / T. 
4. ANOTHER PROOF OF ZARISKI’S LEMMA 315

4. Another proof of Zariski’s Lemma


The following result is a close relative of Lemma 17.16.
Lemma 17.18. Let K be a field and Ω an algebraically closed field. Let S be the
set of all pairs (A, f ) with A is a subring of K and f : A ,→ Ω, partially ordered by
(A, f ) ≤ (A0 , f 0 ) ⇐⇒ A ⊂ A0 and f 0 |A = f.
Then S contains maximal elements, and for any maximal element (B, g), B is a
valuation ring of K.
Proof. An easy Zorn’s Lemma argument shows that S has maximal elements.
Let (B, g) be a maximal element. Put p = Ker(g); since g(B) is a subring of the field
Ω, it is a domain and thus p is a prime ideal of B. By functoriality of localization,
G extends to a homomorphism Bp → Ω. By maximality of (B, g) we have Bp = B,
so that B is a local ring with maximal ideal p. If there existed an element x ∈ K
which is transcendental over the fraction field of B, then B[x] is a polynomial ring
and certainly g extends to B[x]. So K is algebraic over the fraction field of B.
Next let x ∈ K × . We claim that either the ideal pB[x] or pB[x−1 ] is proper.
Indeed this is proved exactly as in Lemma 17.16 above.
Finally, we show that B is a valuation ring of K. Let x ∈ K • . Without loss
of generality, we may assume that pB[x] is a proper ideal of B (otherwise replace
x by x−1 ). Put B 0 = B[x]. By assumption, pB[x] is contained in a maximal ideal
m of B 0 and m ∩ B = p. Hence the embedding of domains B → B 0 induces an
embedding of fields k := B/p ,→ B 0 /m = k 0 . Moreover k 0 is generated over k by
the image of the algebraic element x, so k 0 /k is a finite degree field extension. So g
induces an embedding k ,→ Ω, and since Ω is algebraically closed, this extends to
an embedding k 0 = B 0 /m ,→ Ω. By maximality of B, this implies x ∈ B. 

Remark: It should be possible to consolidate Lemmas 17.16 and 17.18 into a single
result. Let me know if you see how to do it.
Proposition 17.19. Let A ⊂ B be domains with B finitely generated as an
A-algebra. Let β ∈ B • . There exists α ∈ A• satisfying the following property: any
homomorphism f of A into an algebraically closed field Ω with f (α) 6= 0 extends to
a homomorphism f : B → Ω with f (β) 6= 0.
Proof.
Step 0: Induction on the number of generators reduces us to the case B = A[x].
Step 1: Suppose that x is transcendental over A, i.e., B is a univariate polynomial
ring over A. Write
β = an xn + . . . + a1 x + a0 , ai ∈ A
and put α = a0 . If f : A → Ω is such that f (α) 6= 0, then since Ω is infinite, there
exists ζ ∈ Ω such that f (an )ζ n + . . . + f (a1 )ζ + f (a0 ) 6= 0. Using the universal
polynomial of polynomial rings, we may uniquely extend f to a homomorphism
from B to Ω by putting f (x) = ζ, and then f (β) 6= 0.
Step 2: Suppose that x is algebraic over the fraction field of A. Then so is β −1 .
Hence we have equations of the form
an xm + . . . + a1 x + a0 , ai ∈ A
a0m β −m + . . . + a01 β −1 + a00 , a0i ∈ A.
316 17. VALUATION RINGS

Put α = an a0m . Suppose f : A → Ω is any homomorphism with f (α) 6= 0. We


may extend f to a homomorphism from A[α−1 ] → Ω by mapping α−1 to f (α)−1
and then, by Lemma 17.18, to a homomorphism f : C → Ω for some valuation
ring C containing A[α−1 ]. By construction x is integral over A[α−1 ]. Since C is
integrally closed, x ∈ C. Thus C contains B and in particular β ∈ C. Similarly,
β −1 is integral over A[α−1 ] so β −1 ∈ C. Thus β ∈ C × , so f (β) 6= 0. Restricting to
B gives the desired homomorphism. 
Proof of Zariski’s Lemma: Let k be a field and B a field that is finitely generated
as a k-algebra. We want to show that the field extension B/k has finite degree. For
this it is enough to show that B/k is algebraic. In Proposition 17.19 take A = k,
β = 1 and Ω to be an algebraic closure of k. 

5. Discrete valuation rings


5.1. Introducing DVRs.
Proposition 17.20. For a valuation ring R with value group G, the following
are equivalent:
(i) R is a PID.
(ii) R is Noetherian.
(iii) R is an ACCP-domain.
(iv) G≥0 = {x ∈ G | x ≥ 0} is well-ordered.
(v) G is isomorphic to (Z, ≤).
Proof. (i) ⇐⇒ (ii) ⇐⇒ (iii) is a special case of Theorem 16.17.
(iii) ⇐⇒ (iv): a totally ordered set is well-ordered iff there are no infinite strictly
descending chains. But an infinite strictly descending chain in G≥0 gives rise to an
infinite strictly ascending chain of principal ideals in R, and conversely.
(iv) =⇒ (v): First suppose G is Archimedean, so G ,→ R: this endows G with
a topology. If G is discrete, it is generated by its least positive element hence is
order-isomorphic to Z. If G is not discrete, there exists an infinite strictly decreasing
sequence of positive elements of G converging to 0, so G≥0 is not well-ordered. Next
suppose that G is not Archimedean, and choose x, y > 0 such that for all n ∈ Z+ ,
nx < y. Then {y − nx}n∈Z+ is an infinite strictly descending sequence in G≥0 ,
contradicting well-ordering.
(v) =⇒ (iv): famously, the standard ordering on Z≥0 is a well-ordering. 
Exercise 17.20. Show directly: a local PID is a valuation ring with value group
Z.
A valuation ring satisfying the equivalent conditions of Proposition 17.20 is called
a discrete valuation ring (or, sometimes, a DVR).
5.2. Further characterizations of DVRs.

In many ways, discrete valuation rings are – excepting only fields – the simplest
class of rings. Nevertheless they have an important role to play in algebra and
arithmetic and algebraic geometry. One reason for this is as follows: every DVR is
a one-dimensional Noetherian local ring. The converse does not hold.

Example: Let k be a field, and let R be the k-subalgebra of k[t] generated by


5. DISCRETE VALUATION RINGS 317

t2 and t3 . This is a one-dimensional Noetherian domain; the ideal m generated by


t2 and t3 is a nonprincipal maximal ideal. Indeed, even in the localization Rm the
ideal mRm is not principal: consider what its order at 0 would be with respect to
the valuation ordt on k(t): it would have to be 1, but there is no such element in Rm .

The question then is to find necessary and sufficient conditions for a one-dimensional
Noetherian local domain to be a DVR. As we have seen, being a PID is enough,
but again, this is not very useful as whether a one-dimensional domain is a PID is
difficult to check in practice. Remarkably, it turns out if a local, one-dimensional
Noetherian domain has any one of a large number of good properties, it will neces-
sarily be a DVR. Here is the theorem.
Theorem 17.21. (Recognition Theorem for DVRs) Let R be a one-dimensional
Noetherian local domain, with maximal ideal m. the following are equivalent:
(i) R is regular: the dimension of mm2 as an R/m-vector space is 1.
(ii) m is principal.
(iii) R is a PID.
(iv) R is a UFD.
(v) R is integrally closed.
(vi) Every nonzero ideal is of the form mn for some n ∈ N.
Proof. (i) ⇐⇒ (ii): Choose t ∈ π \ π 2 . By assumption, t generates m/m2 ,
so by Nakayama’s Lemma t generates m. Conversely, if m is monogenic as an R-
module, certainly m/m2 is monogenic as an R/m-module.
Evidently (iii) =⇒ (ii). Proposition 16.5 gives (ii) =⇒ (iii) and also (ii) =⇒
(vi). Moreover (iii) ⇐⇒ (iv) by Proposition 16.1 and (iv) =⇒ (v) by 15.14.
Next, for all n ∈ N we have (π)n /(π)n+1 ∼
= R/m, thus R is regular.
(vi) =⇒ (i): Assume that dimR/m m/m2 > 1. Choose u ∈ m \ m2 . Then
m ( hu, m2 i ( m2 .
So we have (i) ⇐⇒ (ii) ⇐⇒ (iii) ⇐⇒ (iv) ⇐⇒ (vi) =⇒ (v).
Finally, we show (v) =⇒ (ii): Let 0 6= x ∈ m. Since m is the only prime ideal
containing (x) we must have r((x)) = m. Since R/(x) is Noetherian, its radical,
m/(x), is nilpotent, so there is a unique least n ∈ Z+ such that mn ⊂ (x). Let
y ∈ mn−1 \ (x) and consider the element q = xy of the fraction field K of R. Since
y∈/ (x), q −1 = xy ∈
/ R; since R is integrally closed in K, q −1 is not integral over R.
Then q−1 m is not contained in m, for otherwise m would be a faithful R[q −1 ]-module
which is finitely generated as an R-module, contradicting Theorem 14.1. But by
construction, q −1 m = xy m ⊂ R, hence q −1 m = R and then m = Rx = (x). 
5.3. Modules over DVRs.
Lemma 17.22. Let R be a DVR with uniformizing element π, and let a ∈ Z+ .
Then the ring Ra = R/(π a ) is self-injective – i.e., Ra is an injective Ra -module.
Exercise 17.21. Prove Lemma 17.22. (Hint: Baer’s Criterion!)
Theorem 17.23. Let R be a DVR with uniformizing element π, and let M be
a nonzero finitely generated R-module.
a) There is N ∈ N and positive integers n, a1 ≥ a2 ≥ . . . ≥ an such that
n
M∼
M
(40) = RN ⊕ R/(π ai ).
i=1
318 17. VALUATION RINGS

b) The numbers N, n, a1 , . . . , an are invariants of the isomorphism class of the mod-


ule M : i.e., they are the same for any two decompositions of M as in (40) above.
Proof.
Step 0: Consider the canonical short exact sequence
0 → M [tors] → M → M/M [tors] → 0.
Since M is a finitely generated module over a Noetherian ring, M [tors] is finitely
generated. Moreover, M/M [tors] is a finitely generated torsionfree module over
a PID, hence is free (Proposition 3.61). Moreover, we know that the rank of a
free module over any (commutative!) ring is well-defined (when R is a domain
with fraction field K, the proof is especially easy: the rank of a free module M is
dimK M ⊗R K), so the invariant N in the statement of the theorem is precisely
the rank of M/M [tors]. Moreover, since M/M [tors] is free – hence projective – the
sequence splits, so
M = RN ⊕ M [tors].
We are reduced to the case of a nonzero finitely generated torsion module M .
Step 1: The annihilator of M is an ideal of R, of which there aren’t so many: it must
be (π a1 ) for some a1 ∈ Z+ . Thus M may be viewed as a faithful Ra1 = R/(π a1 )-
module. Moreover, choosing an element x of M which is not annihilated by π a1 −1 ,
the unique Ra1 -module map Ra1 → M which sends 1 to m is an injection. Taking
M 0 = M/Ra1 , we get a short exact sequence
0 → Ra1 → M → M 0 → 0.
By Lemma 17.22, Ra1 is an injective Ra1 -module, so the sequence splits:
M∼ = Ra ⊕ M 0 .1

Step 2: Since M is finitely generated over Ra1 , it is a quotient of some Artinian


Ra1 -module RaM1 , hence by Theorem 8.4 M is Artinian. Moreover M is a finitely
generated module over the Noetherian ring, so M is also Noetherian. By Theorem
8.14, this means that M has finite length as an R-module. Hence so does its direct
summand M 0 and indeed clearly the length of M 0 is less than the length of M .
This completes the proof of part a) by induction.
Step
Ln 3: So afar we have that a finitely generated torsion R-module is of the form
i=1 R/(π i
) with positive integers a1 ≥ a2 ≥ . . . ≥ an , and with ann(M ) = (π a1 ).
In order to prove the uniqueness statement of part b), it suffices to prove that for
all 0 < b ≤ a, R/(π b ) is an indecomposable R/(π a )-module. If so, then
n
M∼
M
= R/(π ai )
i=1
is simply the decomposition of the finite length module M into indecomposables
described in the Krull-Schmidt Theorem: in particular, since clearly R/(π a ) ∼ =
R/(π b ) implies a = b (consider annihilators), it is unique up to permutation of the
factors. So suppose that R/(π a ) = M1 ⊕ M2 with M1 , M2 nonzero. If π a does
not annihilate M1 , then as above we can find a split embedding R/(π a ) ,→ M1 ,
which contradicts the fact that the length of M1 must be smaller than the length
of R/(π a ). So M1 – and similarly M2 – is annihilated by π a−1 and thus R/(π a )
would be annihilated by π a−1 , a contradiction. 
CHAPTER 18

Normalization theorems

We work in the following situation: R is an integrally closed domain with fraction


field K, L/K is a field extension, and S = IL (R) is the integral closure of R in L.
In more geometric language, S is the normalization of R in the extension L/K.
As above, we may as well assume that L/K is algebraic, since in the general
case, if we let L0 = IK (L) be the algebraic closure of K in L, then S is contained
in L0 anyway. So let us assume this. Then we know that S is integrally closed with
fraction field L. We also know that the Krull dimensions of S and R coincide.

The major questions are the following:

(Q1) Is S finitely generated as an R-module?


(Q2) Is S Noetherian?
(Q3) If not, then can anything nice be said about S?

Note that if R is Noetherian, then an affirmative solution to (Q1) implies an affir-


mative answer to (Q2). Also, the example R = Z, K = Q, L = Q shows that both
(Q1) and (Q2) may have a negative answer if [L : K] is infinite.

1. The First Normalization Theorem


The first, and easiest, result is the following:
Theorem 18.1. (First Normalization Theorem) Let R be an integrally closed
domain with fraction field K, L/K a finite separable field extension, and S =
IL (R).
a) There exists a K-basis x1 , . . . , xn of L such that S is contained in the R-
submodule generated by x1 , . . . , xn .
b) If R is Noetherian, S is a finitely generated R-module.
c) If R is a PID, then S is a free R-module of rank [L : K].
Proof. a) By the proof of Proposition 14.10, for any x ∈ L, there exists
0 6= r ∈ R such that rx ∈ S. Therefore there exists a K-basis u1 , . . . , un of L such
that ui ∈ S for all i.1 Now take x ∈ S and write x = i bi ui with bi ∈ K. Since
P

L/K is separable there are n = [L : K] distinct K-embeddings of L into K, say


2
√ the discriminant ∆ = ∆(u1 , . . . , un ) = (det(σj (ui ))) is nonzero.
σ1 , . . . , σn , and
We may put ∆) = det(σj (ui )). For all 1 ≤ j ≤ n we have
X
σj (x) = bi σj (ui ).
i

1We have not yet used the separability hypothesis, so this much is true in the case of an
arbitrary finite extension.

319
320 18. NORMALIZATION THEOREMS

Using Cramer’s rule, we may solve for the bi to get


√ X X√
∆bi = dij σj (x), dbi = ddij σj (x),
j j

where the dij ’s are polynomials in the σj (ui ) with coefficients in Z. Thus ∆bi and

∆bi are integral over R. Since ∆ ∈ K and R is integrally closed, we have ∆bi ∈ A.
Therefore S is contained in the R-span h u∆1 , . . . , u∆n iR , establishing part a).
b) By part a), S is a submodule of a finitely generated R-module, hence if R is
Noetherian S is finitely generated.
c) We know that S is a submodule of a free rank n R-module; if R is a PID, then S
is a free R-module of rank at most n. Since S ⊗R K = L, the rank must be n. 
This has the following important result, which is the first of three fundamental
finiteness theorems in algebraic number theory, the existence of a finite integral
basis for the ring of integers of any algebraic number field:
Corollary 18.2. Let R = Z, K = Q, L a number field of degree n. Then
the ring ZL = Z ∩ K of all algebraic integers lying in L, is an integrally closed,
Noetherian domain of Krull dimension one which is, as a Z-module, free of rank n.
Proof. Indeed ZL = IL (Z), so by Proposition 14.11, it is integrally closed in
its fraction field L. Since Z is a PID and L/Q is finite separable, Theorem 18.1
applies to show that ZL ∼ = Zn as a Z-module. Being a finitely generated Z-module,
still more is it a finitely generated algebra over the Noetherian ring Z, so it is itself
Noetherian. Since Z, like any PID, has Krull dimension one and ZL is an integral
extension of Z, by Corollary 14.17 ZL also has Krull dimension one. 
A Dedekind domain is a domain which is Noetherian, integrally closed and of
Krull dimension at most one. We will systematically study Dedekind domains in
§20, but for now observe that Corollary 18.2 implies that the ring of integers of an
algebraic number field is a Dedekind domain. In fact, the argument establishes that
the normalization S of any Dedekind domain R in a finite separable field extension
L/K is again a Dedekind domain that is finitely generated as an R-module.

What about the nonseparable case?


Example 18.3. See http: // math. stackexchange. com/ questions/ 1171259
COMPLETE ME! ♣

2. The Second Normalization Theorem


Theorem 18.4. (Second Normalization Theorem) Let R be a domain with frac-
tion field K. Suppose that at least one of the following holds:
• R is absolutely finitely generated – i.e., finitely generated as a Z-algebra – or
• R contains a field k and is finitely generated as a k-algebra.
Let L/K be a finite degree field extension. Then S = IL (R) is a finitely generated
R-module.
Proof. First suppose that R is a finitely generated algebra over a field k.
Step 0: We may assume that L/K is normal. Indeed, let M be the normal closure
of L/K, so M/K is a finite normal extension. Let T be the integral closure of R in
M . If we can show that T is finitely generated over R, then, since R is Noetherian,
the finitely submodule S is also finitely generated over R.
3. THE KRULL-AKIZUKI THEOREM 321

Step 1: We will make use of the field-theoretic fact that if M/K is normal and L is
the maximal purely inseparable subextension of M/K, then M/L is separable [FT,
§6.3]. Let S be the integral closure of R in L and T the integral closure of R in T .
Then T is module-finite over R iff T is module-finite over S and S is module-finite
over R. Suppose we can show that S is module-finite over R. Then S is a finitely
generated R-algebra so S is a Noetherian integreally closed domain, and the module
finiteness of T over S follows from Theorem 18.1. Thus we are redueced to the case
in which L/K is purely inseparable, say [L : K] = q = pa .
Step 2: By Noether Normalization, R is module-finite over a polynomial ring
k[t1 , . . . , td ]. If S is module-finite over k[t1 , . . . , td ], then certainly it is module-
finite over the larger ring R. Thus we may assume without loss of generality that
R = k[t1 , . . . , td ], K = k(t1 , . . . , td ). In particular we may assume that R is inte-
grally closed (in K). For all a ∈ L, NL/K (a) = aq ∈ K. Let k 0 /k be the extension
obtained by adjoining the qth roots of the coefficients of the minimal polynomials of
1/q 1/q
a finite set of generators of L/K, so k 0 /k is finite, so L ⊂ k 0 (t1 , . . . , td ). So it is
1/q 1/q
enough to show that the integral closure of k[t1 , . . . , td ] in k 0 (t1 , . . . , td ) is finite
over k[t1 , . . . , td ]. But in this case the integral closure can be computed exactly: it
1/q 1/q
is k 0 [t1 , . . . , td ] (indeed it is at least this large, and this ring is a UFD, hence
integrally closed), which is finite over k[t1 , . . . , td ]. 

3. The Krull-Akizuki Theorem


In this section we come to one of the most beautiful and useful results in the
subject, the Krull-Akizuki Theorem. Its content is essentially that normalization
works magnificently well in dimension one. Our treatment follows [M, §11].
Lemma 18.5. For a Noetherian domain R, the following are equivalent:
(i) R has dimension at most one.
(ii) For every nonzero ideal I of R, R/I is an Artinian ring.
(iii) For every nonzero ideal I of R, `R (R/I) < ∞.
Proof. (i) =⇒ (ii): R/I is Noetherian, and prime ideals of R/I correspond
to prime ideals of R containing the nonzero ideal I, so are all maximal. By Theorem
8.36, R/I is Artinian.
(ii) =⇒ (iii): Every finitely generated module over an Artinian ring is also Noe-
therian hence has finite length.
¬ (i) =⇒ ¬ (iii): If R has dimension greater than one, there is a nonzero, non-
maximal prime ideal p of R. The R-module R/p is a domain which is not a field,
hence not Artinian, hence of infinite length. 
Lemma 18.6. Let R be a one-dimensional Noetherian domain with fraction field
K. Let M be a torsionfree R-module with r = dimK M ⊗R K < ∞. Then for all
x ∈ R• , `(M/xM ) ≤ r`(R/xR).
Proof. Step 1: First suppose that M is finitely generated. Let η1 , . . . , ηr ∈ M
be R-linearly independent and put E = hη1 , . . . , ηr iM . Since r = dimK M ⊗R K, for
η ∈ M , there is t ∈ R with tη ∈ E. Put C = M/E. Then C is finitely generated, so
there is t ∈ R• such that tC = 0. By Lemma 18.5, the ring R/tR is Artinian. Since
C is a finitely generated R/tR-module, it has finite length, and thus it also has
finite length, say m, as an R-module. For x ∈ R• and n ∈ Z+ , the exact sequence
E/xn E −→ M/xn M → C/xn C
322 18. NORMALIZATION THEOREMS

yields
(41) `(M/xn M ) ≤ `(E/xn E) + `(C).
Since E and M are torsionfree, we have xi M/xi+1 M =∼ M/xM for all i ∈ N and
∼ E/xE; it follows that
similarly xi E/xi+1 E =
n`(M/xM ) ≤ n`(E/xE) + `(C) ∀n ∈ Z+ ,
and thus
`(M/xM ) ≤ `(E/xE).
Since E ∼
= Rr , E/xE ∼
= (R/xR)r , so
`(M/xM ) ≤ `((R/xR)r ) = r`(R/xR).
Step 2: In the general case, put M = M/xM and let N = hω1 , . . . , ωs i be a finitely
generated submodule of M . Lift each ωi to ωi ∈ M and put M1 = hω1 , . . . , ωs i.
We get
`(N ) ≤ `(M1 /M1 ∩ xM ) ≤ `(M1 /xM1 ) ≤ r`(R/xR),
the last inequality by Step 1. Because the right hand side of this inequality is
independent of N , by Exercise 8.13 `(M ) ≤ r`(R/xR).
Step 3: We have `(R/xR) < ∞ by Lemma 18.5. 
Theorem 18.7. (Krull-Akizuki) Let R be a one-dimensional Noetherian do-
main with fraction field K, let L/K be a finite degree field extension of K, and let
S be a ring with R ⊂ S ⊂ L. Then:
a) S is Noetherian of dimension at most 1.
b) If J is a nonzero ideal of S, then S/J is a finite length R-module.
Proof. b) It is no loss of generality to replace L by the fraction field of S. Let
r = [L : K], so that S is a torsionfree R-module of rank r. By Lemma 18.6, for any
x ∈ R• we have `R (S/aS) < ∞. Let J be a nonzero ideal of S and b ∈ J • . Since b
is algebraic over R it satisfies a relation
am bm + . . . + a1 b + a0 = 0, ai ∈ R
of minimal degree. Since S is a domain, a0 ∈ (J ∩ R)• , so
`R (S/J) ≤ `R (S/a0 S) < ∞.
a) Since
`S (J/a0 S) ≤ `R (J/a0 S) ≤ `R (S/a0 S) < ∞,
J/a0 S is a finitely generated S-module. Being an extension of a finitely generated
S-module by a finitely generated S-module, J is itself a finitely generated: S is
Noetherian. If P is a nonzero prime ideal of S then S/P has finite length so is an
Artinian domain, hence a field: S has dimension at most one. 
We remark that S need not be finitely generated as an R-module. Thus Step 2 in
the proof of Lemma 18.6 is actually used in the proof of the Krull-Akizuki Theorem.

Comparing the following exercise with Exercise 8.35 gives a good illustration of
how much simpler things are in dimension 1.
Exercise 18.1. Let k be a field, and let A be a k-subalgebra of k[t]. Show: either
A = k or A is a one-dimensional Noetherian domain that is finitely generated as a
k-algebra.
3. THE KRULL-AKIZUKI THEOREM 323

Corollary 18.8. Let R be a one-dimensional Noetherian domain with fraction


field K, let L/K be a finite degree field extension, and let S be the integral closure
of R in L. Then S is a Dedekind domain, and for every maximal ideal p of R there
are only finitely many prime ideals of S lying over p.
Exercise 18.2. Prove Corollary 18.8.
CHAPTER 19

The Picard Group and the Divisor Class Group

1. Fractional ideals
Let R be a domain with fraction field K. A fractional ideal of R is a nonzero
R-submodule I of K for which there exists 0 6= a ∈ R such that aI ⊂ R – or
equivalently, if I ⊂ a1 R.

When one is talking about fractional R-ideals, one inevitably wants to compare
them to ideals of R in the usual sense, and for this it is convenient to speak of an
integral R-ideal, i.e., an R-submodule of R.
Exercise 19.1. Show: a finitely generated R-submodule of K is a fractional
ideal.
Some texts define a fractional R-ideal to be a finitely generated R-submodule of K,
but this seems wrong because we certainly want every nonzero integral ideal of R
to be a fractional ideal, but if R is not Noetherian then not every integral ideal will
be finitely generated. (It is not such a big deal because most of these references are
interested only in invertible fractional ideals – to be studied shortly – and one of
the first things we will see is that an invertible fractional ideal is necessarily finitely
generated as an R-module.)

We denote the set of all fractional ideals of R by Frac(R).


Theorem 19.1. Let I, J, M be fractional ideals in a domain R.
a) Then
I ∩ J = {x ∈ K | x ∈ I and x ∈ J},
I + J = {x + y | x ∈ I, y ∈ J},
n
X
IJ = { xi yi , | xi ∈ I, yi ∈ J},
i=1
(I : J) = {x ∈ K | xJ ⊂ I}
are all fractional ideals.
b) We may partially order Frac(R) under inclusion. Then the greatest lower bound
of I and J is I ∩ J and the least upper bound of I and J is I + J.
c) If I ⊂ J, then IM ⊂ JM .
d) R itself is a fractional ideal, and R · I = R. Thus the fractional ideals form a
commutative monoid under multiplication.
Proof. a) It is immediate that I ∩J, I +J, IJ and (I : J) are all R-submodules
of K. It remains to be seen that they are nonzero and can be scaled to lie inside
R. Suppose I ⊂ a1 R and J ⊂ 1b R. Then:
325
326 19. THE PICARD GROUP AND THE DIVISOR CLASS GROUP

1
0 ( I ⊂ I + J ⊂ ab R, so I + J is a fractional ideal.
1
0 ( IJ ⊂ I ∩ J ⊂ ab R, so IJ and I ∩ J are fractional ideals.
Since I ∩ R is a fractional ideal, there exists a nonzero c ∈ R lying in I. Then for
y ∈ J, cb y ∈ cR ⊂ I, so cb ∈ (I : J). Similarly, if 0 6= d ∈ J, then ad
1
(I : J) ⊂ R.
Parts b), c) and d) can be easily verified by the reader. 
Proposition 19.2. Let I, J be fractional ideals for a domain R. The map
(I : J) → HomR (J, I), x 7→ (y 7→ xy)
is an isomorphism of R-modules.
Exercise 19.2. Prove it.
Proposition 19.3. All the above operations on fractional ideals commute with
localization: that is, if S ⊂ R• is a multiplicatively closed subset, then
S −1 (I ∩ J) = S −1 I ∩ S −I J,
S −1 (I + J) = S −1 I + S −1 J,
S −1 (IJ) = (S −1 I)(S −1 J),
S −1 (I : J) = (S −1 I : S −1 J).
Exercise 19.3. Prove Proposition 19.3.
A fractional ideal is principal if it is of the form xR for some x = a
b ∈ K •.
Proposition 19.4. For a fractional ideal I of R, the following are equivalent:
(i) I is principal.
(ii) I is monogenic as an R-module.
(iii) I ∼
=R R.
(iv) I is free as an R-module.
Proof. (i) ⇐⇒ (ii): By definition, a monogenic R-module M is one of the
form Rx for some x ∈ M .
(i) =⇒ (iii): For x ∈ K × multiplication by x−1 is an R-module isomorphism from
xR to R.
(iii) =⇒ (ii) and (iii) =⇒ (iv) are immediate.
(iv) =⇒ (iii): It is no loss of generality to assume that I ⊂ R. Since any two
elements x, y ∈ I are R-linearly dependent – indeed (y)x + (−x)y = 0 – if I is free,
it must have rank 1. 
If xR is a principal fractional ideal, so is x−1 R, and we have
(xR)(x−1 R) = R.
Thus, in Frac(R), every principal fractional ideal xR is a unit, with inverse x−1 R.

Let Prin(R) denote the set of all principal fractional ideals of the domain R.
Exercise 19.4. Show: Prin(R) is a subgroup of Frac(R), and we have a short
exact sequence 1 → R× → K × → Prin(R) → 1.
Exercise 19.5. Define the ideal class monoid C(R) = Frac(R)/ Prin(R).
a) Show: C(R) is well-defined as a commutative monoid.
b) Show: C(R)√is trivial iff R is a PID.
c) Show: C(Z[ −3]) is a finite commutative monoid which is not a group.
2. THE IDEAL CLOSURE 327

For a general domain, C(R) need only be a commutative monoid. In the next
section we “repair” this by defining the Picard group Pic(R).

2. The Ideal Closure


For I ∈ Frac(R), put

I ∗ = (R : I) = {x ∈ K | xI ⊂ R}.

Exercise 19.6. Let R be a domain. Show: for any fractional R-ideal I, I ∗ is a


fractional R-ideal. (Hint: if R ⊂ I, then I ∗ ⊂ R. Reduce the general case to this.)

The fractional ideal I ∗ is called1 the quasi-inverse of I. As we shall see later in


this section, if the ideal I has an inverse in the monoid Frac R, then its inverse must
be I ∗ : i.e. II ∗ = R. In general though all we get from the definition of I ∗ is the
relation II ∗ ⊂ R. This observation motivates the following one.

Proposition 19.5. Let R be a domain, and let R ⊂ K × K given by xRy iff


xy ∈ R. Let (Φ, Ψ) be the induced Galois connection from 2K to itself. Then, for
any fractional ideal I of R, Φ(I) = Ψ(I) = I ∗ . In other words, I 7→ I ∗ is a self-dual
antitone Galois connection on Frac R.

Exercise 19.7. Prove Proposition 19.5.

As usual, we denote the associated closure operator by I 7→ I. Now the machinery


of Galois connections gives us many facts for free that we would otherwise have to
spend a little time deriving:

Corollary 19.6. Let R be a domain and let I, J ∈ Frac R.


a) If I ⊂ J, then J ∗ ⊂ I ∗ .
b) We have I ∗ ⊂ J ∗ ⇐⇒ I ⊃ J.
c) We have I = I.
d) We have (I ∩ J)∗ = I ∗ + J ∗ and I ∩ J = I ∩ J.

Exercise 19.8. Prove Corollary 19.6.

Exercise 19.9. Give an example of I, J ∈ Frac R with J ∗ ⊂ I ∗ but I 6⊂ J.

Proposition 19.7. For a domain R and I ∈ Frac R, we have


\
I= d−1 R.
d∈K × | I⊂d−1 R

Proof.

I = (I ∗ )∗ = {x ∈ K |xI ∗ ⊂ R} = {x ∈ K | ∀d ∈ K × , dI ⊂ R =⇒ xd ∈ R}
\
= d−1 R. 
d∈K × | I⊂d−1 R

1Unfortunately?
328 19. THE PICARD GROUP AND THE DIVISOR CLASS GROUP

3. Invertible fractional ideals and the Picard group


Like any monoid, Frac(R) has a group of units, i.e., the subset of invertible elements.
Explicitly, a fractional ideal I is invertible if there exists another fractional ideal
J such that IJ = R. We denote the group Frac(R)× of invertible fractional ideals
by Inv(R).
Exercise 19.10. Let I1 , . . . , In be fractional ideals of R. Show: the product
I1 · · · In is invertible iff each Ii is invertible.
Lemma 19.8. a) For a fractional ideal I, the following are equivalent:
(i) I is invertible.
(ii) II ∗ = R.
b) (To contain is to divide) If I ⊂ J are fractional ideals with J invertible, then
I = J(I : J).
Proof. a) (i) =⇒ (ii): As above, for any fractional ideal I we have II ∗ ⊂ R.
Now suppose there exists some fractional ideal J such that IJ = R, then
J ⊂ (R : I) = I ∗ ,
so
R = IJ ⊂ II ∗ .
(ii) =⇒ (i) is obvious.
b) By definition of (I : K) we have J(I : J) ⊂ I. Conversely, since I ⊂ J, J −1 I ⊂ R.
Since (J −1 I)J = I, it follows that J −1 I ⊂ (I : J) and thus I ⊂ J(I : J). 
Proposition 19.9. Let I be an invertible fractional ideal. Then I is a finitely
generated projective rank one module.
Proof. Step 1: We show an invertible fractional ideal Pn I is a finitely generated
projective module. Since II ∗ = R, we may write 1 = i=1 xi yi with xi ∈ I and
yi ∈ I ∗ . For 1 ≤ i ≤ n, define fi ∈ Hom(I, R) be fi (x) = xyi . Then for all x ∈ I,
X X
x= xxi yi = xi fi (x).
i i
By the Dual Basis Lemma, I is a projective R-module generated by x1 , . . . , xn .
Step 2: To show that I has rank one, it suffices to show that for all p ∈ Spec R, Ip
is free of rank one over Rp . But since projective implies locally free, we know that
Ip is a free Rp -module of some rank, and for any ring R and any ideal I, I cannot
be free of rank greater than one over R. Indeed, if so I would have two R-linearly
independent elements x and y, which is absurd, since yx + (−x)y = 0. 
Conversely:
Proposition 19.10. Let I be a nonzero fractional ideal of R which is, as an
R-module, projective. Then I is invertible.
Proof. We have the inclusion ι : II ∗ ⊂ R which we wish to show is an equality.
This can be checked locally: i.e., it is enough to show that for all p ∈ Spec R,
ιp : Ip Ip∗ → Rp is an isomorphism. By Proposition 19.3, it is equivalent to show
that Ip (Ip )∗ → Rp is an isomorphism, but since I is projective, by Kaplansky’s
Theorem Ip is free. As above, being a nonzero ideal, it is then necessarily free
of rank one, i.e., a principal fractional ideal xRp . It follows immediately that
(Ip )∗ = x−1 Rp and thus that the map is an isomorphism. 
3. INVERTIBLE FRACTIONAL IDEALS AND THE PICARD GROUP 329

To sum up:
Theorem 19.11. Let R be a domain, and let I be a fractional R-ideal. Then
I is invertible iff it is projective, in which case it is projective of rank one.
For any R-module M , the R-dual is defined to be M ∨ = Hom(M, R). There is a
canonical R-bilinear map T : M ∨ × M → R obtained by mapping (f, x) 7→ f (x).
This induces an R-linear map T : M ∨ ⊗R M → R. We say that an R-module M is
invertible if T is an isomorphism.
Proposition 19.12. Consider the following conditions on an R-module M .
(i) M is rank one projective.
(ii) M is invertible.
(iii) There exists an R-module N and an isomorphism T : M ⊗R N ∼ = R.
Then (i) =⇒ (ii) =⇒ (iii) always, and (iii) =⇒ (i) if M is finitely generated.
Proof. (i) =⇒ (ii): We have a map T : M ∨ ⊗ M → R so that it suffices to
check locally that is an isomorphism, but M is locally free so this is easy.
(ii) =⇒ (iii) is immediate.
(iii) =⇒ (i): Since M is finitely generated, by Theorem 13.35 to show that M is
projective it is enough to show that for all p ∈ Spec R Mp is free of rank one. Thus
we may as well assume that (R, m) is a local ring with residue field R/m = k. The
base change of the isomorphism T to R/m is an isomorphism (recall that tensor
product commutes with base change)
Tk : M/mM ⊗k N/mN → k.
This shows that dimk M/mM = dimk N/mN = 1, so in particular M/mM is
monogenic as an R/m-module. By Nakayama’s Lemma the lift to R of any generator
x of M/mM is a generator of M , so M is a monogenic R-module and is thus
isomorphic to R/I for some ideal I. But indeed I = ann(M ) ⊂ ann(M ⊗R N ) =
ann(R) = 0, so M ∼
= R/(0) = R is free of rank one. 
Theorem 19.13. (Cohen) Let R be a domain.
a) The set of invertible ideals of R is an Oka family in the sense of § 4.5.
b) If every nonzero prime ideal of R is invertible, then every nonzero fractional
ideal of R is invertible.
Proof. a) Let I ⊂ J be ideals of R with J and (I : J) invertible (this implies
I 6= 0). By Lemma 19.8, I = J(I : J) and thus, as the product of two invertible
ideals, I is invertible. Since for any ideals I, J of R we have (I : J) = (I : I + J),
by taking J = hI, xi for any x ∈ R we recover the Oka condition.
b) Seeking a contradiction, suppose I is a nonzero ideal of R which is not invertible.
Consider the partially ordered set S of ideals containing I which are not invertible.
Then the union of any chain in S is a non-invertible ideal: indeed, if it were invertible
then by Proposition 19.9 it would be finitely generated and thus equal to some
element in the chain: contradiction. Thus by Zorn’s Lemma there is a nonzero
ideal J which is maximal element in the family of ideals which are not invertible.
By part a) and the Prime Ideal Principle, J is prime: contradiction. 
Theorem 19.14. Let I and J be invertible fractional ideals. Then there is a
canonical isomorphism of R-modules

I ⊗R J → IJ.
330 19. THE PICARD GROUP AND THE DIVISOR CLASS GROUP

Proof. The natural multiplication map I × J → IJ is R-bilinear so factors


through an R-module map m : I ⊗R J → IJ. Once we have a globally defined map,
to see that it is an isomorphism it is enough to check it locally: for all p ∈ Spec R,

mp : Ip ⊗Rp Jp → Ip Jp
and we are thus allowed to assume that I and J are principal fractional ideals. This
makes things very easy, and we leave the endgame to the reader. 
Corollary 19.15. Let I and J be invertible fractional R-ideals. the following
are equivalent:
(i) There exists x ∈ K × such that xI = J.
(ii) I ∼
=R J, i.e., I and J are isomorphic R-modules.
Proof. (i) =⇒ (ii): If J = xI, then multiplication by x gives an R-module
isomorphism from I to J.
(ii) =⇒ (i): Since I ∼
=R J we have
I J = I ⊗R J ∼
−1 ∼ −1
= I −1 ⊗R I ∼
= II −1 = R.
By Proposition 19.4, I −1 J is a principal fractional ideal, i.e., there exists x ∈ K ×
such that I −1 J = xR. Multiplying through by I, we get xI = J. 
Proposition 19.16. Let M be a rank one projective module over a domain R
with fraction field K. Then there is a fractional R-ideal I such that M ∼
=R I.
Proof. Since M is projective, it is flat, and so tensoring the injection R ,→ K
with M we get an injection f : M = R⊗R M ,→ M ⊗R K ∼ = K, the last isomorphism

since M is locally free of rank 1. Thus f : M → f (M ), and f (M ) is a finitely
generated R- submodule of K and thus a fractional R-ideal. 
Putting together all the pieces we get the following important result.
Theorem 19.17. Let R be a domain. The following two commutative groups
are canonically isomorphic:
(i) Inv(R)/ Prin(R) with [I][J] := [IJ].
(ii) Isomorphism classes of rank one projective R-modules under tensor product.
We may therefore define the Picard group Pic R to be either the group of invertible
fractional ideals modulo principal fractional ideals under multiplication or the group
of isomorphism classes of rank one projective modules under tensor product.
Lemma 19.18. In any domain R, let P1 , . . . , Pk be a set of invertible prime
ideals and let Q1 , . . . , Ql be any set of prime ideals. Suppose that
k
Y l
Y
Pi = Qj .
i=1 j=1

Then i = j and there exists some permutation σ of the set {1, . . . , k} such that for
all 1 ≤ i ≤ k we have Pi = Qσ(i) .
In other words, prime factorization is unique for products of invertible primes.
Proof. Assume
Q without loss of generality that P1 does not strictly contain
any Pi . Since j Qj ⊂ P1 , some Qj , say Q1 , is contained in P1 . Similarly, since
Q
i Pi ⊂ Q1 , there exists i such that Pi ⊂ Q1 . Thus Pi ⊂ Q1 ⊂ P1 . By our
assumption on the minimality of P1 , we have P1 = Pi = Q1 . We can thus cancel
P1 = Q1 and obtain the result by induction. 
3. INVERTIBLE FRACTIONAL IDEALS AND THE PICARD GROUP 331

Lemma 19.19. Let R be an integrally closed Noetherian domain with fraction


field K, and let I be a fractional R-ideal. Then (I : I) := {x ∈ K | xI ⊂ I} = R.
Proof. Clearly R ⊂ (I : I). Conversely, let x ∈ (I : I). Then I is a faithful
R[x]-module that is finitely generated over R, so x is integral over R. 

Lemma 19.20. Let R be a domain with fraction field K, S ⊂ R \ {0} a multi-


plicative subset, and I, J fractional R-ideals.
a) We have (I + J)S = IS + JS .
b) (IJ)S = IS JS .
c) (I ∩ J)S = IS ∩ JS .
d) If I is finitely generated, then (I ∗ )S = (IS )∗ .
Proof. Parts a) and b) are immediate and are just recorded for future refer-
ence. For part c), we evidently have (I ∩ J)S ⊂ IS ∩ JS . Conversely, let x ∈ IS ∩ JS ,
so x = si1 = sj2 with i ∈ I, j ∈ J, s1 , s2 ∈ S. Put b = a1 s2 = a2 s1 ∈ I ∩ J; then
x = s1bs2 ∈ (I ∩ J)S , establishing part c). For part d), note first that (I + J)∗ =
I ∗ ∩ J ∗ . Also if 0 6= x ∈ K, then (Rx)
TnS = RS x. Hence if I = Rx1 + . . . + Rx
Tnn , then
IS = RS x1 +. . .+RS xn , so (IS )∗ = i=1 x1i RS . On the other hand, I ∗ = i=1 x1i R,
and thus part c) we have
n
\ 1
(I ∗ )S = RS = (IS )∗ . 
i=1
x i

Lemma 19.21. A nonzero ideal in a Noetherian domain contains a product of


nonzero prime ideals.
Proof. Assume not: let I be a nonzero ideal which is maximal with respect
to the property of not containing a product of nonzero prime ideals. Then I is not
prime: there are x1 , x2 ∈ R \ I such that x1 x2 ∈ I. For i = 1, 2 put Ii := hI, xi i,
so that I ( Ii and I ⊃ I1 I2 . By maximality of I, I1 ⊃ p1 · · · pr and I2 ⊃ q1 · · · qs
(with pi , qj prime for all i, j), and then I ⊃ p1 · · · pr q1 ⊃ qs , contradiction. 

Lemma 19.22. (Jacobson) Let R be a Noetherian domain of Krull dimension


at most one. Let I be a proper, nonzero ideal of R. Then (R : I) ) R.
Proof. Let 0 6= a ∈ I, so aR ⊂ I ⊂ R. By Lemma 19.21, there are nonzero
prime ideals p1 , . . . , pm such that aR ⊃ p1 · · · pm ; we may assume m is minimal. Let
m be a maximal ideal containing I. Then m ⊃ I ⊃ aR ⊃ p1 · · · pm ; since nonzero
prime ideals are maximal, this implies m = pi for some i, say for i = 1. If m = 1
then I = aR so (R : I) = a−1 R ) R. Now suppose m > 1; by minimality of m, aR
does not contain p2 · · · pm so we may choose b ∈ p2 · · · pm \ aR. Put c = a−1 b. Then
c∈/ R and cI ⊂ cm = a−1 bm ⊂ amp2 · · · pm ⊂ a−1 (aR) = R, so c ∈ (R : I). 

The following result gives information about when a prime ideal is invertible.
Proposition 19.23. Let R be a Noetherian domain, and p a nonzero prime
ideal of R. If p is invertible, then it has height one and Rp is a DVR.
Proof. Since p is invertible, Rp is a Noetherian local domain with a principal
maximal ideal pRp . By Theorem 17.21, Rp is a DVR, and thus p has height one. 
332 19. THE PICARD GROUP AND THE DIVISOR CLASS GROUP

4. Divisorial ideals and the Divisor Class Group


For I, J ∈ Frac(R), we write I ≤ J if every principal fractional ideal Ra which is
contained in I is also contained in J. This gives a preordering on Frac(R). Let R
be the associated equivalence relation, i.e., I ∼ J if I ≤ J and J ≤ I.

We write D(R) = Frac(R)/ ∼. Elements of D(R) are called divisors on R. For


a fractional ideal I, we denote its image in D(R) by div(I), and for a principal
fractional ideal aR, we write simply div(a). Such elements are called principal
divisors. A fractional ideal I is divisorial if I = I.
Exercise 19.11. Let R be a domain and I a fractional R-ideal.
a) Show: I is the unique divisorial ideal with div I = div I.
b) Show: I is the smallest divisorial fractional ideal containing I.
c) Show: invertible fractional ideals are divisorial.
Exercise 19.12. If I is a divisorial fractional ideal and x ∈ K • , then Ix is a
divisorial fractional ideal.
Exercise
T 19.13. Let {Ii } be a family of divisorial fractional ideals such that
I= i Ii is nonzero. Show: I is a divisorial fractional ideal.

Lemma 19.24. Let I, J ∈ Frac R.


a) The following are equivalent:
(i) I ≤ J.
(ii) J ∗ ⊂ I ∗ .
b) The following are equivalent:
(i) I ∼ J.
(ii) I ∗ = J ∗ .
(iii) I = J.
(iv) div I = div J.
Exercise 19.14. Prove Lemma 19.24.
Proposition 19.25. For I, J, M ∈ Frac R: div I ≤ div J =⇒ div IM ≤
div JM .
Proof. By hypothesis J ⊂ I; equivalently (R : I) = I ∗ ⊂ J ∗ = (R : J). Then
(IM )∗ = (R : IM ) = ((R : I) : M ) ⊂ ((R : J) : M ) = [R : JM ) = (JM )∗ .
It follows that JM ⊂ IM , so div IM ≤ div JM . 
Proposition 19.26. Let R be a domain.
a) For div I, div J ∈ D(R), the operation div I + div J = div IJ is well-defined and
endows D(R) with the structure of a commutative monid.
b) (D(R), +, <) is a lattice-ordered monoid.
Proof. We have
(R : IJ) = ((R : I) : J) = (R : I) : J) = ((R : J) : I)
= ((R : J) : I) = (R : IJ),
so div I + div J = div I + div J.
b) For I, J, M ∈ Frac R with div I ≤ div J, by Proposition 19.25 and part a),
div I + div M = div IM ≤ div JM = div J + div M,
4. DIVISORIAL IDEALS AND THE DIVISOR CLASS GROUP 333

and thus the partial ordering is compatible with the monoid structure. To show
that we have a lattice, for any I, J ∈ Frac R, we need to find the supremum and
infimum of div I and div J. We claim that in fact we have
div(I ∩ J) = sup div I, div J
div(I + J) = inf div I, div J.
To see this we may assume I and J are divisorial. By Exercise 19.12, I ∩ J is
divisorial, so it is clear that for any divisorial ideal M ,
(div I ≤ M, div J ≤ M ) ⇐⇒ (M ⊂ I, M ⊂ J)
⇐⇒ (M ⊂ I ∩ J) ⇐⇒ div I ∩ J ≤ div M.
Next, observe that since I, J ⊂ I + J, div I + J ≤ div I, div J, i.e., div I + J is
a lower bound for {div I, div J}. Conversely, if M ∈ Frac R is such that div M ≤
div I, div J, then I, J ⊂ M so I + J ⊂ M and I + J ⊂ M = M and div M =
div M ≤ div I + J = div I + J. 
Tournant Dangereux One may wonder why we work with divisors at all since
every divisor is represented by a unique divisorial ideal. However, if I and J are
divisorial fractional ideals, the product IJ need not be divisorial.
Proposition 19.27.
a) In D(R) every nonempty set which is bounded above admits a least upper bound.
Explicitly, if (Ii ) is a nonempty
T family of fractional ideals which is bounded above,
then supi (div Ii ) = div( i I i ).
b) In D(R) every nonempty set which is bounded below admits a greatest lower
bound. Explicitly, if (Ji ) is a nonempty
P family of fractional ideals which is bounded
below, then inf i (div Ji ) = div( i Ii ).
c) D(R) is a lattice.
Proof. Bourbaki, p. 477. COMPLETE ME! ♣ 
Theorem 19.28. For a domain R, the following are equivalent:
(i) D(R) is a group.
(ii) R is completely integrally closed.
Proof. (i) =⇒ (ii): Let x ∈ K × . Suppose there is d ∈ R• such that dxn ∈ R
for all n ∈ Z+ . Then I = hR, aiR ∈ Frac R and aI ⊂ I. Then
div I ≤ div aI = div a + div I.
Since D(R) is a group, div R = 0 ≤ div a, and since aR and R are divisorial, a ∈ R.
(ii) =⇒ (i): We’ll show: for all divisorial fractional ideals I, (II ∗ )∗ = R∗ = R,
hence div I + div I ∗ = div R = 0. By Proposition 19.7, it’s enough to show that II ∗
and R are contained in the same principal fractional ideals. Since II ∗ ⊂ R, any
principal fractional ideal which contains R contains II ∗ . Thus, let x ∈ K × be such
that II ∗ ⊂ xR; we want to show R ⊂ xR, i.e., x−1 ∈ R. Suppose that for y ∈ K ×
we have I ⊂ yR, so y −1 ∈ I ∗ and thus Iy −1 ⊂ xR; equivalently, x−1 I ⊂ yR. Thus
x−1 I is contained in every principal fractional ideal containing I, so x−1 I ⊂ I = I.
It follows that x−n I ⊂ I for all n ∈ Z+ . Let w ∈ R• be such that wI ⊂ R. Then
dx−n I ⊂ R, and if z ∈ I • then (wz)x−n ∈ R for all n ∈ Z+ . Since dc ∈ R• and R
is completely integrally closed, by Theorem 14.38 x−1 ∈ R. 
334 19. THE PICARD GROUP AND THE DIVISOR CLASS GROUP

Let P (R) be the image in D(R) of the principal fractional ideals. Then P (R) is a
subgroup of D(R). Thus if R is completely integrally closed (e.g. Noetherian and
integrally closed!) we may form the quotient
Cl R = D(R)/P (R),
the divisor class group of R.
Exercise 19.15. Let R be a completely integrally closed domain. Show: there
is a canonical injection
Pic R ,→ Cl R.
Theorem 19.29. Let R = C[x, y, z]/(xy − z 2 ). Then R is a Noetherian inte-
grally closed domain with Pic R = 0 and Cl R ∼
= Z/2Z.
CHAPTER 20

Dedekind domains

A Dedekind domain is a domain which is Noetherian, integrally closed, and of


dimension at most one. A Dedekind domain has dimension zero iff it is a field.
Although we endeavor for complete precision here (why not?), the reader should
be warned that in many treatments the zero-dimensional case is ignored, when
convenient, in statements of results.

1. Characterization in terms of invertibility of ideals


Theorem 20.1. For a domain R, the following are equivalent:
(i) R is Dedekind: Noetherian, integrally closed of dimension at most one.
(ii) Every fractional R-ideal is invertible.
(iii) Every nonzero prime ideal of R is invertible.
Proof. (i) =⇒ (ii): Let R be a Noetherian, integrally closed domain of
dimension at most one, and let I be a fractional R-ideal. Then II ∗ ⊂ R and hence
also II ∗ (II ∗ )∗ ⊂ R, so I ∗ (II)∗ ⊂ I ∗ . It follows from Lemma 19.19 that (II ∗ )∗ ⊂ R;
moreover, since II ∗ ⊂ R, Lemma 19.22 implies II ∗ = R, i.e., I is invertible.
(ii) =⇒ (i): Since invertible ideals are finitely generated, if every nonzero ideal
is invertible, then R is Noetherian. Let p be a nonzero, nonmaximal prime ideal
of R, so that there exists a maximal ideal m which 0 ( p ( m. By the mantra
“to contain is to divide” for invertible fractional ideals, there exists some invertible
integral ideal I such that p = mI. Suppose that I ⊂ p. Then I = RI ⊃ mI = p, so
we would have p = I and then m = R, contradiction. Then there are x ∈ m \ p and
y ∈ I \ p such that xy ∈ p, contradicting the primality of p.
Finally, we check that R is integrally closed: let x = cb be a nonzero element of
K which is integral over R, so there exist a0 , . . . , an−1 ∈ R such that
xn + an−1 xn−1 + . . . + a1 x + a0 = 0.
Let M be the R-submodule of K generated by 1, x, . . . , xn−1 ; since M is finitely
generated, it is a fractional R-ideal. We have M 2 = M , and thus – since M is
invertible – M = R. It follows that x ∈ R.
(ii) =⇒ (iii) is immediate.
(iii) =⇒ (ii) by Theorem 19.13. 
Recall that a ring R is hereditary if every ideal of R is a projective R-module.
Corollary 20.2. A domain R is hereditary iff it is a Dedekind domain.
Proof. By Theorem 20.1 a domain R is a Dedekind domain iff every fractional
ideal of R is invertible, and clearly the latter condition holds iff every nonzero
integral ideal of R is invertible. Moreover, by Theorem 19.11, a nonzero ideal of a
ring is invertible iff it is projective as an R-module. 
335
336 20. DEDEKIND DOMAINS

2. Ideal factorization in Dedekind domains


Here we will show that in a Dedekind domain every nonzero integral ideal factors
uniquely into a product of primes and derive consequences for the group of invertible
ideals and the Picard group. (The fact that factorization – unique or otherwise! –
into products of primes implies invertibility of all fractional ideals – is more delicate
and will be pursued later.)
Lemma 20.3. Let I be an ideal in a ring R. If there exist J1 , J2 ideals of R,
each strictly containing I, such that I = J1 J2 , then I is not prime.
Proof. Choose, for i = 1, 2, xi ∈ Ji \ I; then x1 x2 ∈ I, so I is not prime. 

Theorem 20.4. Every proper integral ideal in a Dedekind domain has a unique
factorization into a product of of prime ideals.
Proof. After Lemma 19.18 it suffices to show that a nonzero proper integral
ideal I in a Dedekind domain R factors into a product of primes. Suppose not,
so the set of ideals which do not so factor is nonempty, and (as usual!) let I be a
maximal element of this set. Then I is not prime, so in particular is not maximal:
let p be a maximal ideal strictly containing I, so I = pJ. Then J = p−1 I strictly
contains I so factors into a product of primes, hence I does. 

If I is any nonzero integral ideal of I and p is any nonzero prime ideal of a Dedekind
domain R, then we may define ordp (I) via the prime factorization
Y
I= pordp (I) .
p

The product extends formally over all primes, but as I is divisible by only finitely
many primes, all but finitely many exponents are zero, so it is really a finite product.
Corollary 20.5. Let R be a Dedekind domain.
a) The monoid M(R) of nonzero integral ideals is a free commutative monoid on
the set of nonzero prime ideals.
b) The fractional ideals form a free commutative group on the set of prime ideals:
M
Frac(R) = Z.
06=p ∈ Spec R

Proof. Part a) is simply the statement of unique factorization into prime


elements in any commutative monoid. In the group I(R) of all fractional ideals,
the subgroup G generated by the nonzero primes is a free commutative group on
the primes: this just asserts that for primes p1 , . . . , pr and integers n1 , . . . , nr , the
equation pn1 1 · · · pnr r = R implies n1 = . . . = nr = 0, which is easily seen – e.g. by
localizing. Since any fractional ideal J is of the form x1 I with I an integral ideal,
decomposing I and (x) into their prime factorizations expresses J as a Z-linear
combination of prime ideals, so Frac(R) = G. 

Corollary 20.5 allows us to extend the definition of ordp to any fractional R-ideal.

Since for a Dedekind domain there is no distinction between invertible fractional


ideals and all fractional ideals, the Picard group takes an especially simple form: it
is the quotient of the free commutative group Frac(R) of all fractional ideals modulo
3. LOCAL CHARACTERIZATION OF DEDEKIND DOMAINS 337

the subgroup Prin(R) = K × /R× of principal fractional ideals. We therefore have


a short exact sequence
0 → Prin(R) → Frac(R) → Pic(R) → 0,
and also a slightly longer exact sequence
0 → R× → K × → Frac(R) → Pic(R) → 0.
Theorem 20.6. For a Dedekind domain R, the following are equivalent:
(i) Pic(R) = 0.
(ii) R is a PID.
(iii) R is a UFD.
(iv) The set of nonprincipal prime ideals is finite.
Proof. Evidently each fractional ideal is principal iff each integral ideal is
principal: (i) ≡ (ii). Since R has dimension at most one, (ii) ⇐⇒ (iii) by
Proposition 16.1. Evidently (ii) =⇒ (iv), so the interesting implication is that
(iv) implies the other conditions. So assume that the set of (nonzero) nonprincipal
prime ideals is nonempty but finite, and enumerate them: p1 , . . . , pn . Let I be an
integral ideal, and suppose that
I = pa1 1 · · · pann qb11 · · · qbmm .
(As usual, we allow zero exponents.) By the Chinese Remainder Theorem we may
choose an α ∈ R such that ordpi (α) = ai for all i.1 Now consider the fractional
ideal (α−1 )I; it factors as
(α−1 )I = qb11 · · · qbmm rc11 · · · rcl l ,
where the ri ’s are some other prime ideals, i.e., disjoint from the pi ’s. But all of
the (fractional) ideals in the factorization of (α−1 )I are principal, so (α−1 )I = (β)
for some β ∈ K × and then I = (αβ) is principal! 
Exercise 20.1. a) Consider the ring

R1 = Z[ −3] = Z[t]/(t2 + 3).
Show: R1 is a one-dimensional√ Noetherian
√ domain with exactly one nonprincipal
prime ideal, namely p2 = h1 + −3, 1 − −3i.
b) For any n ∈ Z+ , exhibit a ring Rn which is one-dimensional Noetherian and has
exactly n nonprincipal prime ideals.

3. Local characterization of Dedekind domains


Theorem 20.7. Let R be a domain.
a) If R is Dedekind and S is a multiplicative subset, then S −1 R is Dedekind.
b) If R is a Dedekind domain and 0 6= p is a prime ideal of R, then Rp is a DVR.
Proof. Being Noetherian, dimension at most one and integrally closed are
all preserved under localization, so part a) is immediate. Similarly, if 0 6= p is a
prime ideal, then the localization Rp is a local, one-dimensional integrally closed
Noetherian domain, hence by Theorem 17.21 a DVR, establishing b). 
1Note that we want equality, not just ord (α) ≥ a , so you should definitely think about
Pi i
how to get this from CRT if you’ve never seen such an argument before.
338 20. DEDEKIND DOMAINS

Exercise 20.2. Let R be Dedekind with fraction field K; let 0 6= p ∈ Spec R.


a) Show: the map ordp : K × → Z defined above is nothing else than the discrete
valuation corresponding to the localization Rp .
b) Conversely, let v : K × → Z be a discrete valuation. Show that the valuation ring
Rv = v −1 (N) is the localization of R at some maximal ideal p.

4. Factorization into primes implies Dedekind


Theorem 20.8. (Matusita [Ma44]) Let R be a domain with the property that
every nonzero proper integral ideal is a product of prime ideals. Then R is Dedekind.
Proof. Step 1: Let p be an invertible prime of R. We show that p is maximal.
Let a ∈ R \ p, and suppose that ha, pi ( R. Let us then write
I1 := ha, pi = p1 · · · pm ,

I2 := ha2 , pi = q1 · · · qn ,
where the pi and qj are prime ideals. By assumption, I1 ) p, and, since p is prime,
we have also I2 ) p. Therefore each pi and qj strictly contains p. In the quotient
R = R/p we have
(a) = aR = p1 · · · pm
and
(a2 ) = a2 R = q1 · · · qn .
The principal ideals (a) and (a2 ) are invertible, and the pi and qj remain prime in
the quotient. Therefore, we have
q1 · · · qn = p21 · · · pm 2 .
Thus the multisets {{q1 , . . . , qn } and {p1 , p1 , . . . , pm , pm }} coincide, and pulling
back to R the same holds without the bars. Thus
I12 = ha, pi2 = p21 · · · p2m = q1 · · · qn = ha2 , pi,
so
p ⊂ ha, pi2 = a2 R + ap + p2 ⊂ aR + p2 .
So if p ∈ p, p = ax + y with x ∈ R, y ∈ p2 , so ax ∈ p, and since a ∈ R \ p, x ∈ p.
Thus p ⊂ ap + p2 ⊂ p, so p = ap + p2 . Multiplication by p−1 gives R = a + p,
contrary to hypothesis. So p is maximal.
Step 2: Let p be any nonzero prime ideal in R, and 0 6= b ∈ p. Then p ⊃ bR
and
bR = p1 · · · pm ,
with each pi invertible and prime. Thus by Step 1 the pi ’s are maximal. Since
p is prime we have p ⊃ pi for some i and then by maximality p = pi , hence p is
invertible. Since by assumption every proper integral ideal is a product of primes,
we conclude that every integral ideal is invertible, which, by Theorem 20.1 implies
that R is Dedekind. 

Let a and b be ideals of a domain R. We say that b divides a if there is an ideal c


such that bc = a.
5. GENERATION OF IDEALS IN DEDEKIND DOMAINS 339

Exercise 20.3. Suppose a, b are ideals of a domain R such that b divides a.


a) Show: b ⊃ a.
b) Show: c ⊂ (a : b).
c) Can we have c ( (a : b)?

Proposition 20.9. For a Noetherian domain R, the following are equivalent:


(i) R is a Dedekind domain.
(ii) To contain is to divide: For all ideals a, b of R, b ⊃ a ⇐⇒ b divides a.
Proof. (i) =⇒ (ii): The statement is trivial if b = (0). Otherwise, b is
invertible so a = b(a : b) by Lemma 19.8.
(ii) =⇒ (i): We claim that every proper nonzero ideal of R is a product of prime
ideals. Since R is Noetherian, if this is not the case there is an ideal a which is
maximal with respect to not having this property. Let p be a maximal ideal with
a ⊂ p. By hypothesis, there is an ideal c with a = p1 c. Then c ⊃ a. Suppose we
had equality; then repeatedly substituting a = p1 a Tgives a = pk1 a for all k ∈ Z+ ,

and then by the Krull Intersection Theorem, a ⊂ k=1 pk1 = (0), contradiction.
So c properly contains a, so we may write c = p2 · · · pr and thus a = p1 p2 · · · pr :
contradiction. 

Theorem 20.10. For a domain R which is not a field, the following are equiv-
alent:
(i) R is Noetherian, integrally closed, and of Krull dimension one.
(ii) Every fractional (equivalently, every integral) R-ideal is invertible.
(iii) R is Noetherian, and the localization at every maximal ideal is a DVR.
(iv) Every nonzero proper integral ideal factors into a product of prime ideals.
(iv0 ) Every nonzero proper integral ideal factors uniquely into a product of primes.
(v) R is Noetherian, and to contain is to divide for all ideals of R.

5. Generation of ideals in Dedekind domains


Theorem 20.11. Let R be a Dedekind domain and I a nonzero ideal of R.
Then the quotient ring R/I is a principal Artinian ring.
Qr
Proof. Write I = i=1 pai i . By the Chinese Remainder Theorem,
r
R/I ∼
Y
= R/pai i .
i=1

Each factor R/pai iis also a quotient of the localized ring Rp /pai i , which shows that
it is Artinian and principal. Finally, a finite product of Artinian (resp. principal
ideal rings) remains Artinian (resp. a principal ideal ring). 

This has the following striking consequence:


Theorem 20.12. (Asano-Jensen) For a domain R, the following are equiva-
lent:
(i) R is a Dedekind domain.
(ii) For any nonzero ideal I of R and any nonzero element a ∈ I, there exists b ∈ I
such that I = ha, bi.
340 20. DEDEKIND DOMAINS

Proof. The direction (i) =⇒ (ii) follows immediately from Theorem 20.11.
Conversely, assume condition (ii) holds. By Theorem 20.10 it suffices to show that
R is Noetherian and that its localization at each nonzero prime ideal p is a DVR.
Certainly condition (ii) implies Noetherianity; moreover it continues to hold for
nonzero ideals in any localization. So let I be a nonzero ideal in the Noetherian
local domain (Rp , p). It follows that there exists b ∈ p such that p = Ip + bRp . By
Nakayama’s Lemma, I = bRp , so Rp is a local PID, hence a DVR. 
Proposition 20.13. ([J2, Ex. 10.2.11]) Let R be a Dedekind domain, I a
fractional ideal of R and J a nonzero integral ideal of R. Then there exists a ∈ I
such that aI −1 + J = R.
Proof. Let p1 , . . . , ps be the prime ideals of R dividing J. For each 1 ≤ i ≤ r,
choose ai ∈ Ip1 · · · pr p−1
i \ Ip1 · · · pr . Put a = a1 + . . . + ar . We claim that
aI −1 + J = R. It is enough to check this locally. For every prime q 6= pi , we have
JRq = Rq . On the other hand, for all 1 ≤ i ≤ r, aI −1 is not contained in pi , so its
pushforward to Rpi is all of Rpi . 

6. Modules over a Dedekind domain


The main aim of this section is to prove the following important result.
Theorem 20.14. Let M be a finitely generated module over a Dedekind domain.
a) P = M/M [tors] is a finitely generated projective R-module, say of rank r.
b) If r = 0 then of course M = M [tors]. If r ≥ 1 then

M∼ = M [tors] ⊕ P ∼
= M [tors] ⊕ Rr−1 ⊕ I,
with I a nonzero ideal of R.
c) The class [I] of I in Pic R is an invariant of M .
d) There exists N ∈ Z+ , maximal ideals pi and positive integers ni such that
N
M [tors] ∼
M
= R/pni i .
i=1

Much of the content of the main theorem of this section lies in the following converse
of Proposition 3.8b) for finitely generated modules over a Dedekind domain.
Theorem 20.15. For a finitely generated module M over a Dedekind domain,
the following are equivalent:
(i) M is projective.
(ii) M is flat.
(iii) M is torsionfree.
Proof. Of course (i) =⇒ (ii) =⇒ (iii) for modules over any domain, and
we have seen that (i) ≡ (ii) for finitely generated modules over a Noetherian ring.
So it suffices to show (iii) =⇒ (i).
Suppose R is a Dedekind domain and M is a finitely generated nonzero tor-
sionfree R-module. By Proposition 3.8c), we may assume that M ⊂ Rn for some
n ≥ 1. We prove the result by induction on n. If n = 1, then M is nothing else
than a nonzero ideal of R, hence invertible by Theorem 20.10 and thus a rank one
projective module by Theorem 19.11. So we may assume that n > 1 and that every
finitely generated torsionfree submodule of Rn−1 is projective. Let Rn−1 ⊂ Rn be
6. MODULES OVER A DEDEKIND DOMAIN 341

the span of the first n − 1 standard basis elements. Let πn : Rn → R be projection


onto the nth factor, and consider the restriction of πn to M :
π
0 → M ∩ Rn−1 → M →
n
πn (M ) → 0.
Put I = πn (M ). Then I is an ideal of R, hence projective, so the sequence splits:
M → (M ∩ Rn−1 ) ⊕ I.
Now M ∩ Rn−1 is a torsionfree, finitely generated (since M is finitely generated and
R is Noetherian) submodule of Rn−1 , hence is projective by induction. Certainly
a direct sum of projective modules is projective, so we’re done. 
The method of proof immediately yields the following important corollary:
Corollary 20.16. Let P be a finitely generated rank r projective module
Lr over
a Dedekind domain R. Then we have a direct sum decomposition P ∼ = i=1 Ii ,
where each Ii is a nonzero rank one projective R-module.
Let M a finitely generated module over the Dedekind domain R. We have:
0 → M [tors] → M → M/M [tors] → 0.
Put P := M/M [tors]. Then P is finitely generated and torsionfree by Proposition
3.8a), hence projective (by Theorem 20.15), and the sequence splits:
M∼= M [tors] ⊕ P.
Lemma 20.17.
Ln I1 , . . . , In be fractional ideals in the Dedekind domain R. Then
the R-modules i=1 Ii and Rn−1 ⊕ I1 · · · In are isomorphic.
Proof. We will prove the result when n = 2. The general case follows by an
easy induction argument left to the reader.
Choose 0 6= a1 ∈ I1 . Applying Proposition 20.13 with I = I2 and J = a1 I1−1 ⊂ R,
that there exists a2 ∈ I2 such that a1 I1−1 + a2 I2−1 = R. That is there exist bi ∈ Ii−1
such that a1 b1 + a2 b2 = 1. The matrix
 
b1 −a2
b2 a1
is invertible with inverse  
−1 a1 a2
A = .
−b2 b1
For (x1 , x2 ) ∈ I1 ⊕ I2 , we have
y1 = x1 b1 + x2 ∈ R, y2 = −x1 a2 + x2 a1 ∈ I1 I2 .
On the other hand, if y1 ∈ R and y2 = c1 c2 ∈ I1 I2 , then
x1 = a1 y1 − b2 c1 c2 ∈ I1 , x2 = a2 y1 + b1 c1 c2 ∈ I2 .
Thus [x1 x2 ] 7→ [x1 x2 ]A gives an R-module isomorphism from I1 ⊕I2 to R⊕I1 I2 . 
Thus we may write
r
M = M [tors] ⊕ M/M [tors] ∼
M
= M [tors] ⊕ Ii = M [tors] ⊕ Rr−1 ⊕ (I1 · · · Ir ),
i=1
which establishes Theorem 20.14a).

As for part b) of the theorem, let T be a finitely generated torsion R-module.


342 20. DEDEKIND DOMAINS

We notice that the statement of the classification is identical to that of finitely


generated torsion modules over a PID. This is no accident, as we can easily re-
duce to the case of a PID – and indeed to that of a DVR, which we have already
proven (Theorem 17.23). Namely, let I beQthe annihilator of T , and (assuming
r ai
T 6= 0, asQwe certainly may) write I = i=1 pi . Then T is a module over
∼ r ri ∼ L r ai
R/I = R/ i=1 Lpri = i=1 R/pi . By Exercise X.X in §3.1, T naturally decom-
poses as T = i=1 Ti , where Ti is a module over R/pai i . This gives the primary
decomposition of T . Moreover, each Ti is a module over the DVR Rp , so Theorem
17.23 applies.
Corollary 20.18. For any Dedekind domain R, the Picard group Pic R is
^
canonically isomorphic to the reduced K0 -group K0 (R).

Proof. Let P be a finitely generated projective R-module of rank r ≥ 1.


According to Theorem 20.14c) the monoid of isomorphism classes of finitely gen-
erated projective R-modules is cancellative: this means that the canonical map
ϕ : Pic(R) → K0 (R) is injective. It follows easily that the composite map Φ :
ϕ
Pic(R) → K0 (R) → K ^ ^
0 (R) is an injection: indeed, for ϕ(I) to be killed in K0 (R)
but not K0 (R) it would have to be a fractional ideal which has rank zero as an
R-module, and there are no such things. Now an arbitrary nonzero finitely gener-
ated projective R-module is isomorphic to Rr−1 ⊕ I, hence becomes equal to the
class of the rank one module I in K ^0 (R), so Φ is surjective. To check that it is a
homomorphism of groups we may look on a set of generators – namely, the classes
of rank one projective modules. Let us use [P ] for the class of the projective module
P in K0 (R) and [[P ]] for its image in K ^0 (R). Then by Lemma 20.17 we have

Φ([I1 ⊗ I2 ]) = [[I1 ⊗ I2 ]] = [[I1 I2 ]] = [[R ⊕ I1 I2 ]] = [[I1 ⊕ I2 ]] = [[I1 ]] + [[I2 ]]. 


Exercise 20.4. State and prove an appropriate analogue of Proposition 6.13
for finitely generated projective modules over a Dedekind domain R.
Corollary 20.19. For a module M over a Dedekind domain, the following
are equivalent:
(i) The module M is torsionfree.
(ii) The module M is flat.
Proof. (i) =⇒ (ii): If M is finitely generated and torsionfree, then M is flat
by Theorem 20.15. By Corollary 3.93, every torsionfree R-module is flat.
(ii) =⇒ (i): This holds for modules over any domain: Proposition 3.36. 
Corollary 20.20. For a Noetherian domain R, the following are equivalent:
(i) Every torsionfree R-module is flat.
(ii) The ring R is a Dedekind domain.
Proof. (i) =⇒ (ii): Let I be an ideal of R. Then I is a torsionfree R-module,
hence flat by assumption. Since R is Noetherian, the ideal I is finitely generated,
hence projective by Corollary 7.31. By Theorem 19.11, I is invertible. In particular
every nonzero prime ideal of R is invertible, so by Theorem 20.1.
(ii) =⇒ (i): This is part of Corollary 20.19. 
The full characterization of domains in which every torsionfree module is flat is
coming up soon: Theorem 21.10.
7. INJECTIVE MODULES 343

7. Injective Modules
Theorem 20.21. For a domain R with fraction field K, the following are equiv-
alent:
(i) R is Dedekind.
(ii) Every divisible R-module is injective.
Proof. (i) =⇒ (ii): Let D be a divisible R-module. We will show D is
injective using Baer’s Criterion: let I be an ideal of R and f : I → D a module
map. We may assume that I is nonzero and thus, since R is a Dedekind domain,
invertible:Pnif I = ha1 , . . . , an i, there are b1 , . . . , bn ∈ K such that bi I ⊂ R for all i
and 1 = i=1 ai bi . Since D is divisible, there are d1 , . . . , dn ∈ D with f (ai ) = ai di
for all i. Then for x ∈ I,
X X X X
f (x) = f ( bi ai x) = (bi x)f (ai ) = (bi x)ai di = x (bi ai )ei .
i i i
Pn
Put d = i=1 (bi ai )di . Thus F : R → D by x 7→ dx lifts f .
(ii) =⇒ (i): Let I be injective. Then I is divisible and a quotient of a divisible
module is divisible, so every quotient of I is divisible, and thus by assumption every
quotient of I is injective. By Corollaries 3.58 and 20.2, R is Dedekind. 
As an application, we will prove a generalization to Dedekind domains of a non-
trivial result in commutative group theory. Given an commutative group A, it is
natural to ask when its torsion subgroup A[tors] is a direct summand of A, so that
A is the direct sum of a torsion group and a torsionfree group. It is easy to see
that this happens when A is finitely generated, because then A/A[tors] is a finitely
generated torsionfree module over a PID, hence projective. The following exercise
shows that some condition is necessary.
Q
Exercise 20.5. Let A = p Z/pZ, where the product extends over all prime
numbers. Show that A[tors] is not a direct summand of A.
These considerations should serve to motivate the following result.
Theorem 20.22. Let M be a module over a Dedekind domain R. If M [tors] =
M [r] for some r ∈ R, then M [tors] is a direct summand of M .
Proof. Step 1: We claim that if A is a torsionfree R-module, then for every
R-module N , Ext1R (M, N ) is divisible.
proof of claim Let V = A ⊗R K. Since A is torsionfree, we have an exact
sequence
0 → A → V → V /A → 0.
Applying the cofunctor Hom(·, B), a portion of the long exact Ext sequence is
Ext1R (V, B) → Ext1R (A, B) → Ext2R (V /A, B).
Since R is hereditary, by Proposition X.X, Ext2R (A, B) = 0 so Ext1R (A, B) is a quo-
tient of Ext1R (V, B). Since V is a K-module, so is Ext1R (V, B) and thus Ext1R (V, B)
and its quotient Ext1R (A, B) is a divisible module, hence injective by Theorem 20.21.
Step 2: Let T = M [tors] = M [r]. We will show that the sequence
0 → T → M → M/T → 0
splits by computing Ext1R (M/T, T )
= 0. Since M/T is torsionfree, by Step 1
Ext1R (M/T, T ) is divisible. On the other hand, since T = T [r], Ext1R (M/T, T ) =
344 20. DEDEKIND DOMAINS

Ext1R (M/T, T )[r]. Thus multiplication by r on Ext1R (M/T, T ) is on the one hand
surjective and on the other hand identically zero, so Ext1R (M/T, T ) = 0. By Theo-
rem 3.89 the sequence splits. 
CHAPTER 21

Prüfer domains

A Prüfer domain is a domain in which each finitely generated ideal is invertible.1


Exercise 21.1. Show: every Bézout domain is a Prüfer domain.

1. Characterizations of Prüfer Domains


One might be forgiven for thinking the invertibility of finitely generated ideals is a
somewhat abstruse condition on a domain. The following result shows that, on the
contrary, this determines a very natural class of domains.
Theorem 21.1. (Characterization of Prüfer Domains) For a domain R, the
following are equivalent:
(i) R is a Prüfer domain: every nonzero finitely generated ideal is invertible.
(i0 ) Every nonzero ideal of R generated by two elements is invertible.
(ii) Nonzero finitely generated ideals are cancellable: if a, b, c are ideals of R and a
is finitely generated and nonzero, then ab = ac =⇒ b = c.
(iii) For every p ∈ Spec R, Rp is a valuation ring.
(iii0 ) For every m ∈ MaxSpec R, Rm is a valuation ring.
(v) For all ideals A, B, C of R, A(B ∩ C) = AB ∩ AC.
(vi) For all ideals A, B of R, (A + B)(A ∩ B) = AB.
(vii) If A and C are ideals of R with C finitely generated and A ⊂ C, then there
exists an ideal B of R such that A = BC.
(viii) For all ideals A, B, C of R with C finitely generated, we have
(A + B : C) = (A : C) + (B : C).
(ix) For all ideals A, B, C of R with C finitely generated, we have
(C : A ∩ B) = (C : A) + (C : B).
(x) For all ideals A, B, C of R, A ∩ (B + C) = A ∩ B + A ∩ C.
Proof. We will show: (i) ⇐⇒ (i0 ),
(i) ⇐⇒ (ii), (i) =⇒ (iii) =⇒ (iv) =⇒ (v) =⇒ (vi) =⇒ (ii), (i) =⇒
(vii) =⇒ (iv), (iv) =⇒ (viiii) =⇒ (ii), (iv) =⇒ (ix) =⇒ (ii), and (iv) ⇐⇒
(x). This suffices!
(i) =⇒ (i0 ) is immediate.
(i0 ) =⇒ (i): We go by induction on the number of generators. A nonzero ideal
with a single generator is principal, hence invertible. By assumption, every nonzero
ideal generated by two elements is invertible. Hence we may assume that n ≥ 3

1Except for the zero ideal, of course.

345
346 21. PRÜFER DOMAINS

and that every nonzero ideal of R generated by n − 1 elements is invertible, and let
c = hc1 , . . . , cn i. We may assume ci 6= 0 for all i. Put
a = hc1 , . . . , cn−1 i, b = hc2 , . . . , cn i,
d = hc1 , cn i, e = c1 a−1 d−1 + cn b−1 d−1 .
Then
ce = (a + hcn i)c1 a−1 d−1 + (hc1 i + b)cn b−1 d−1
= c1 d−1 + c1 cn a−1 d−1 + c1 cn b−1 d−1 + cn d−1
= c1 d−1 (R + cn b−1 ) + cn d−1 (R + c1 a−1 ).
−1 −1
Since cn b , c1 a ⊂ R, we get
ce = c1 d−1 + cn d−1 = hc1 , cn id−1 = R.
(iii) =⇒ (iii0 ) is immediate. (iii0 ) =⇒ (iii): if p ∈ Spec R, let m be a maximal
ideal containing p. Then Rp is an overring of Rm , and every overring of a valuation
ring is a valuation ring.
(i) =⇒ (ii) is immediate, since invertible ideals are cancellable.
(ii) =⇒ (iii): First suppose a is a nonzero finitely generated ideal and b, c are
ideals of R with ab ⊂ ac. Then ac = ab + ac = a(b + c); cancelling a gives c = b + c,
so b ⊂ c. Now let p be a prime ideal of R. By Exercise 17.6, it is enough to show
that for any as , bt ∈ Rp , we have either ( as ) ⊂ ( bt ) or ( bt ) ⊂ ( as ). Since 1s , 1t ∈ R× , it
is equivalent to show that (a) ⊂ (b) or (b) ⊂ (a): for this we may assume a, b 6= 0.

Theorem 21.2. Let R be a domain.
a) Suppose R is a GCD-domain. Then R is Prüfer iff it is Bézout.
b) A Prüfer UFD is a PID.
Proof. a) Since principal ideals are invertible, any Bézout domain is a Prüfer
domain. Conversely, suppose R is a GCD-domain and a Prüfer domain. Let x, y ∈
R• and let d be a GCD of x, y. Certainly we have (d) ⊃ hx, yi. Thus ι : hx, yi ,→ (d)
is a homomorphism of R-modules which we want to show is an isomorphism. By
the Local-Global Principle for Module Homomorphisms it is enough to show that
for all p ∈ Spec R, ιp is an isomorphism of Rp -modules, i.e., hx, yiRp = hdiRp .
By Proposition 15.16, d is again the GCD of x and y in the valuation ring Rp
(equivalently, the valuation of d is the minimum of the valuations of x and y) so
that the principal ideal hx, yiRp is generated by hdiRp .
b) Suppose R is a Prüfer UFD. By part a) R is Bézout, and by Theorem 16.17 a
Bézout UFD is a PID. 
Proposition 21.3. For a Prüfer domain R, the following are equivalent:
(i) R is a Bézout domain.
(ii) Pic(R) = 0.
Proof. In the Prüfer domain R, an ideal I is invertible iff it is finitely gener-
ated. So (i) and (ii) each assert that every finitely generated ideal is principal. 
Proposition 21.4. A Prüfer domain is integrally closed.
Proof. In Theorem 20.1 we showed that a domain in which all fractional R-
ideals are invertible is integrally closed. In the proof we only used the invertbility of
finitely generated fractional ideals, so the argument works in any Prüfer domain. 
1. CHARACTERIZATIONS OF PRÜFER DOMAINS 347

Exercise 21.2. Prove Corollary 21.4 using the local nature of integral closure.
1.1. A Chinese Remainder Theorem for Prüfer domains.

Recall that we have a Chinese Remainder Theorem which is valid in any ring:
Theorem 4.20. There is however another useful version of the Chinese Remainder
Theorem which holds in a domain R iff R is a Prüfer domain.

Let R be a ring, let I1 , . . . , In be a finite sequence of ideals in R and let x1 , . . . , xn


be a finite sequence of elements in R. We may ask: when is there an element x ∈ R
such that x ≡ xi (mod Ii ) for all i?
If we assume the ideals Ii are pairwise comaximal, then this holds in any ring
by CRT (Theorem 4.20). But suppose we drop that condition. Then, if such an x
exists, we have x − xi ∈ Ii for all i, hence for all i and j,
(42) xi − xj = (x − xj ) − (x − xi ) ∈ Ii + Ij .
Thus we get a necessary condition (which, notice, is vacuous when the ideals are
pairwise comaximal). Let us say that a ring has property ECRT(n) if for all
ideals I1 , . . . , In and elements x1 , . . . , xn satisfying (42), there exists x ∈ R such
that x ≡ xi (mod Ii ) for all i. We say that R satisfies ECRT (Elementwise Chinese
Remainder Theorem) if it satisfies ECRT(n) for all n ∈ Z+ .
Exercise 21.3. Show: a PID satisfies property ECRT.
Lemma 21.5. Any ring satisfies ECRT(1) and ECRT(2).
Proof. ECRT(1) is trivial. For ECRT(2): let I, J be ideals of R, let x1 , x2 ∈
R, and suppose x1 − x2 ∈ I + J: there are i ∈ I, j ∈ J such that x1 − x2 = i + j.
Put x = x1 − i = x2 + j. Then x ≡ x1 (mod I) and x ≡ x2 (mod J). 
Theorem 21.6. For a ring R, the following are equivalent:
(i) ECRT holds in R.
(ii) ECRT(3) holds in R.
(iii) For all ideals A, B, C in R, A + (B ∩ C) = (A + B) ∩ (A + C).
(iv) For all ideals A, B, C in R, A ∩ (B + C) = (A ∩ B) + (A ∩ C).
Proof. (i) =⇒ (ii) is immediate.
(ii) =⇒ (iii): The inclusion A + (B ∩ C) ⊂ (A + B) ∩ (A + C) holds for ideals in
any ring. Conversely, let t ∈ (A + B) ∩ (A + C). Then by ECRT(3) there is x ∈ R
satisfying all of the congruences
x≡0 (mod A),
x ≡ t (mod B),
x ≡ t (mod C),
and thus x ∈ A, x − t ∈ B ∩ C, so t = x − (x − t) ∈ A + (B ∩ C).
(iii) =⇒ (iv): For A, B, C ideals of R, we we have
(A ∩ B) + (A ∩ C) = ((A ∩ B) + A) ∩ ((A ∩ B) + C) = A ∩ ((A ∩ B) + C)
and
(A ∩ B) + (A ∩ C) = (A + (A ∩ C)) ∩ ((A ∩ C) + B) = A ∩ ((A ∩ C) + B),
and thus
(A ∩ B) + C = (A ∩ C) + B.
348 21. PRÜFER DOMAINS

It follows that
(A ∩ B) + C = (A ∩ B) + C + (A ∩ C) + B = B + C
and thus
(A ∩ B) + (A ∩ C) = A ∩ ((A ∩ B) + C) = A ∩ (B + C).
(iv) =⇒ (iii): Assume (iv). Then for all ideals A, B, C of R,
(A + B) ∩ (A + C) = (A + B) ∩ A + (A + B) ∩ C = A ∩ (A + B) + C ∩ (A + B)
= (A∩A)+(A∩B)+(A∩C)+(B∩C) = A+(A∩B)+(A∩C)+(B∩C) = A+(B∩C).
(iii) =⇒ (i): We go by induction on n. Having established that ECRT(1) and
ECRT(2) hold in any ring, we let n ≥ 2, assume ECRT(n) and show ECRT(n + 1):
let x1 , . . . , xn+1 ∈ R and I1 , . . . , In+1 be ideals of R such that xi − xj ∈ Ii + Ij for
all 1 ≤ i, j ≤ n. By ECRT(n), there Tnis y ∈ R with y ≡ xi (mod I)i for 1 ≤ i ≤ n.
We claim that y − xn+1 ∈ In+1 + i=1 Ii .
proof of claim: Since we have assumed (iii), we have by induction that
n
\ n
\
a+ bi = (a + bi ),
i=1 i=1
and in particular
n
\ n
\
In+1 + Ii = (Ii + In+1 ).
i=1 i=1
Also, for all 1 ≤ i ≤ n, we have
y − xn+1 = (y − xi ) + (xi − xn+1 ) ∈ Ii + Ii + In+1 ∈ Ii + In+1
and thus indeed
n
\ n
\
y − xn+1 ∈ (Ii + In+1 ) = In+1 + Ii .
i=1 i=1
Because of the claim and ECRT(2), there is t ∈ R satisfying
n
\
t≡y (mod Ii ),
i=1

t ≡ xn+1 (mod In+1 ).


Then for 1 ≤ i ≤ n,
t − xi = (t − y) + (y − xi ) ∈ Ii .


2. Butts’s Criterion for a Dedekind Domain


One of our first results on Dedekind domains was Theorem 20.4: in a Dedekind
domain –defined as a Noetherian, integrally closed domain of dimension at most
one – every nonzero ideal factors uniquely as a product of prime ideals. It was then
natural to ask about the converse: if in a domain R every nonzero proper ideal is
uniquely a product of prime ideals, must R be Dedekind? We proved a result of
Matusita which is stronger than this: a domain in which every nonzero proper ideal
is a product of prime ideals is necessarily a Dedekind domain: uniqueness of the
product was not required.
2. BUTTS’S CRITERION FOR A DEDEKIND DOMAIN 349

In particular if all ideals factor into primes then all ideals factor uniquely into
primes. Let’s try to show this directly: suppose we have nonzero prime ideals
p1 , . . . , pr , q1 , . . . , qs in a domain such that
p1 · · · pr = q1 · · · qs .
Then p1 ⊃ q1 · · · qs , so by X.X, p1 ⊃ qj for some j. It is natural to try to deduce
p1 = qj from this. This certainly holds if each qj is maximal, so we get the desired
implication when R has dimension one. In general it seems that we have acquired
nothing more than further respect for Matusita’s Theorem, but the above idea will
be used in the proof of the main result of this section.

The deduction p1 ⊃ qj =⇒ p1 = qj also holds if p1 = (p) and qj = (q) are


both principal: for principal ideals, to contain is to divide, so we get q = xp, and
then because prime elements are irreducible elements we have x ∈ R× so p1 = qj .
Neither of the above steps works in general. By Proposition X.X, a Noetherian
domain in which to contain is to divide is necessarily a Dedekind domain. General-
izing irreducible elements to ideals we get the condition on a nonzero proper ideal
a that a = bc implies b = R or c = R. We cannot call such an a “irreducible” since
that is already taken for a condition involving intersections of ideals, so following
H. Butts we call such an ideal unfactorable. An ideal a is factorable if it is not
unfactorable, i.e., if there are proper ideals b and c such that a = bc.

Example: Let R be a valuation ring with valuation group Q and maximal ideal
m. Then m = m · m. Thus it is possible for a nonzero prime (even maximal) ideal
to be factorable. This also shows that it is possible for an ideal to be a product of
prime ideals in more than one way.

As usual in factorization theory, the ascending chain condition makes things nicer.
Exercise 21.4. Let R be a Noetherian domain.
a) Show: every nonzero ideal of R is a (finite, of course) product of unfactorable
ideals. (Products over sets of cardinality 0 and 1 are allowed.)
b) Show that a nonzero prime ideal of R is unfactorable.
A domain in which each nonzero nonunit factors uniquely into a product of irre-
ducible elements is, by definition, a UFD. So it is natural to ask in which domains
each nonzero proper ideal factors uniquely into a product of unfactorable ideals. A
theorem of H. Butts gives the answer.
Theorem 21.7. (Butts [Bu64]) Let R be a domain in which each nonzero
proper ideal factors uniquely as a product of unfactorable ideals. Then R is Dedekind.
Proof. Step 1: Unique factorization into unfactorables implies: for ideals a, b,
if ac = bc then a = b. By Theorem 21.1, R is Prüfer.
Step 2: We will show that every nonzero prime ideal of R is invertible and apply
Theorem 20.1. So let p be a nonzero prime ideal, and let p ∈ p• . Let (p) = u1 · · · ur
be the unique factorization of p into unfactorable ideals. Since (p) is invertible, so
is each ui ; by Proposition 19.9, each ui is finitely generated. Let x ∈ R \ ui . Then
Ui = ui + hxi is a finitely generated ideal in a Prüfer domain, hence invertible,
so by Lemma 19.8 ui = Ui (ui : Ui ). Since ui is unfactorable, either Ui = R or
350 21. PRÜFER DOMAINS

(ui : Ui ) = R. If (ui : Ui ) = R then ui = Ui , contradiction. So Ui = R. Thus ui is


maximal. Since p ⊃ u1 · · · ur , we have p = ui for some i, so p is invertible. 

3. Modules over a Prüfer domain


Recall that a module is semihereditary if every finitely generated submodule is
projective and that a ring R is semihereditary if the module R is semihereditary:
i.e., every finitely generated ideal of R is projective.
Proposition 21.8. A domain R is a semihereditary iff it is a Prüfer domain.
Exercise 21.5. Prove Proposition 21.8.
Lemma 21.9. Let R be a domain, and let M be a finitely generated torsionfree
R-module. Then M is a submodule of a finitely generated free module.
Proof. Since M is torsionfree, M ,→ M ⊗R K, and ι : M ⊗R K ∼ = K n for
some n ∈ N. Since M is finitely generated, there exists x ∈ R• such that the image
of xM in M ⊗R K is contained in Rn , and thus ι ◦ (x•) : M ,→ Rn . 
Theorem 21.10. For a domain R, the following are equivalent:
(i) Every torsionfree R-module is flat.
(ii) Every finitely generated torsionfree R-module is projective.
(iii) R is a Prüfer domain.
Proof. (i) =⇒ (ii): Let M be a finitely generated torsionfree R-module. By
assumption M is flat, and since R is a domain, by Corollary 13.36 M is projective.
(ii) =⇒ (iii): Finitely generated ideals are assumed projective, hence invertible.
(iii) =⇒ (i): Let R be a Prüfer domain and M a torsionfree R-module. Then
M = limi Mi is the direct limit of its finitely generated submodules, hence a direct
−→
limit of finitely generated torsionfree modules Mi . By Lemma 21.9, each Mi is a
finitely generated submodule of a free R-module. By Theorems 21.8 and 3.66, each
Mi is projective, hence flat. Thus M is a direct limit of flat modules, hence is itself
a flat module by Corollary 3.92. 
CHAPTER 22

One Dimensional Noetherian Domains

1. Residually Finite Domains


For an ideal I in a ring R, we put |I| = #R/I (a possibly infinite cardinal).

R is a residually finite ring if or every nonzero ideal I, R/I is finite.


Exercise 22.1. Show: finite rings and fields are residually finite.
Exercise 22.2. Let R be a ring such that R/(a) is finite for all a ∈ R• . Show:
R is residually finite.
Lemma 22.1. Let 0 ≤ I ⊂ J be ideals in a residually finite ring. Then:
a) |J| | |I|.
b) We have I = J ⇐⇒ |I| = |J|.
Exercise 22.3. Prove it.
Proposition 22.2. Let R be a residually finite ring that is not a field. Then:
a) R is Noetherian.
b) We have dim R ≤ 1.
c) The following are equivalent:
(i) R is a domain.
(ii) dim R = 1.
d) The following are equivalent:
(i) R is finite.
(ii) dim R = 0.
Proof. a) Since for any nonzero ideal I, R/I is finite, we can say even more:
the length of a chain of ideals starting with I is at most the number of prime divisors
(counted with multiplicity) of |I|.
b) If p is a nonzero prime ideal of R, then R/p is a finite domain, hence a field, so
p is maximal. Thus dim R ≤ 1.
c) If R is a domain, then – since it is not a field – dim R ≥ 1. Combining with part
b) we get dim R = 1. Inversely, if R is not a domain, then (0) is not prime, so for
every prime ideal p of R, R/p is finite, hence p is maximal: dim R = 0.
d) Clearly if R is finite then it has dimension 0. Conversely, if dim R = 0 then by
part a) R is Artinian. By Theorem 8.35 we may reduce to the case in which R is
Artinian local with nilpotent maximal ideal: suppose e is the least positive integer
such that me = 0. Since R is not a field, e > 1 and thus R/m is a finite field. Then
for all i ∈ N, mi /mi+1 is a finitely generated module over the finite field R/m, so
it’s finite. Thus #R = #R/m#m/m2 · · · #me−1 /me < ∞. 
Theorem 22.3. (Samuel [Sa71]) Let R be a Noetherian ring, and let n ∈ Z+ .
Then the set of ideals I of R with |I| = n is finite.
351
352 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

Proof. Since the number of isomorphism classes of rings of cardinality n is


finite, it is enough to fix any ring S of cardinality n and show that the set {bi }i∈I
of ideals of R such ∼
T that R/bi = S is finite.
Putting b = i∈I bi , we have a monomorphism of rings
R/bi ∼
Y
(43) B := R/b ,→ = SI .
i∈I

Let m1 , . . . , mr be the maximal ideals of the (finite, hence Artinian) ring S, and for
Tr 1 ≤ j ≤ r, let qj be the cardinality of the finite field S/mj . Then m1 · · · mr =
each
j=1 mj is the Jacobson radical which coincides with the nilradical, hence there
exists s ∈ Z+ such that (m1 · · · mr )s = 0. Let P (t) ∈ Z[t] be the polynomial
r
Y
P (t) = (tqj − t)s .
j=1

Then for any x ∈ S and any 1 ≤ j ≤ r, xqj − x ∈ mj , so P (x) = 0. It follows that


for all X = (xi ) ∈ S I , P (X) = (P (xi )) = 0. From (43) it follows that P (x) = 0
for all x ∈ B. Since the nonzero polynomial P has only finitely many roots in any
domain, for each prime ideal p of B, we conclude that B/p is finite. Thus B is
Noetherian of Krull dimension 0 hence is Artinian. But by the structure theory for
Artinian rings, any Artinian ring with finite residue fields is actually finite (cf. the
proof of Proposition 22.2d)). That is, R/b is finite, so there are only finitely many
ideals of R containing b. In particular I is finite. 
Proposition 22.4. Let (0) 6= I, J be ideals of the residually finite domain R.
a) If I and J are comaximal – i.e., I + J = R – then |IJ| = |I||J|.
b) If I is invertible, then |IJ| = |I||J|.
Proof. Part a) follows immediately from the Chinese Remainder Theorem.
As for part b), we claim that the norm can be computed locally: for each p ∈ ΣR ,
let |I|p be the norm of the ideal IRp in the local finite norm domain Rp . Then
Y
(44) |I| = |I|p .
p
Tn
To see this, let I = i=1 qi be a primary decomposition of I, with pi = rad(qi ). It
follows that {q1 , . . . , qn } is a finite set of pairwise comaximal ideals, so the Chinese
Remainder Theorem applies to give
n
R/I ∼
Y
= R/qi .
i=1

Since R/qi is a local ring with maximal ideal corresponding to pi , it follows that
|qi | = |qi Rpi |, establishing the claim.
Using the claim reduces us to the local case, so that we may assume the ideal
I = (xR) is principal. In this case the short exact sequence of R-modules
xR R R
0→ → → →0
xJ xJ (x)J
together with the isomorphism
R ·x xR

J xJ
does the job. 
1. RESIDUALLY FINITE DOMAINS 353

Theorem 22.5. (Butts-Wade) For a residually finite domain R, the following


are equivalent:
(i) R is a Dedekind domain.
(ii) For all nonzero ideals a, b of R, |ab| = |a||b|.
Proof. (i) =⇒ (ii): Apply Theorem 20.1 and Proposition 22.4.
(ii) =⇒ (i): Step 1: By 22.2, R is Noetherian. So if we can show that to contain
is to divide, Theorem X.X applies to show that R is Dedekind. Let a ⊂ b be ideals
of R. The divisibility trivially holds if a = (0), so we may assume that a 6= (0).
Step 2: Suppose that we can show that if b = a + hbi we have
a = b(a : b).
Then the general case follows: since R is Noetherian we may write b = a +
hb1 , . . . , bn i. For 1 ≤ i ≤ n put bi = a + hb1 , . . . , bi i. By Step 3,
a = b1 (a : b1 ), b1 = b2 (b1 : b2 ), . . . , b = bn−1 (b : bn−1 ),
and thus
a = (a : b1 )(b1 : b2 ) · · · (bn−1 : b)b.
Step 3: So suppose b = a + hbi. Certainly b(a : b) ⊂ a. Since (a : b) ⊃ a, (a : b)
and hence also b(a : b) is not zero, so the above containment gives
|a| | |b(a : b)| = |b||(a : b)|.
|a|
Since a ⊂ b, k = |b| ∈ Z+ and
|a|
| |(a : b)|.
|b|
So if we can show |(a : b)| ≤ k, then |a| = |b||(a : b)| = |b(a : b)|, and by Lemma
22.1, b(a : b) = a. Let {xi }ki=1 be a set of coset representatives for a in b, and let
{yj }nj=1 be a set of coset representatives for (a : b) in R. We will define an injection
from the first set to the second, which suffices to complete the proof. For 1 ≤ i ≤ n,
byi ∈ b, so there is a unique 1 ≤ i ≤ k such that byj ∈ a + xi , and we define
Φ : {yj }nj=1 → {xi }ki=1 , yj 7→ xi .
(If y ∈ R is such that y + (a : b) = yj + (a : b), then b(y − yj ) ∈ b(a : b) ⊂ a, so
by + a = byj + a: that is, the mapping is well-defined on cosets, independent of the
chosen representatives. But that is not necessary for the argument to go through.)
We check the injectivity: if 1 ≤ u, v ≤ n are such that byu , byv ∈ a + xj , then
b(yu − yv ) ∈ a so yu − yv ∈ ((b) : a) = ((b) + a, a) = (b, a), so u = v. 

Exercise 22.4. Let R be a residually finite domain. By Theorem 22.3, for all
n ∈ Z+ , there are only finitely many ideals I with N (I) ≤ n. We can therefore
define a formal Dirichlet series
X
ζR (s) = N (I)−s ,
I){0}

the Dedekind zeta function of


Q R.
a) Show that ζZ (s) = n n1s = p 1−p1 −s = p ζZ(p) (s), where the products extend
P Q

over all prime numbers p.


354 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

b) Let R = Fp [t]. Show that ζR (s) = 1−p11−s .


c) Suppose R is a Dedekind domain. Show that
Y
ζR (s) = ζRp (s).
p∈MaxSpec R

Exercise 22.5. Let R = Z[t]/(t2 + 3) (or, equivalently, Z[ −3]). Let q = (2).
a) Show: there is a unique ideal p2 with R/p2 = Z/2Z. Evidently p2 is maximal.
b) Show: r(q) = p2 , and deduce that I is primary.
c) Show: q is not a prime power, and indeed, cannot be expressed as a product of
prime ideals.1

2. Cohen-Kaplansky domains
Let R be a domain. For a cardinal number κ, the statement “R has κ irreducibles”
will mean that there is a subset P ⊂ R such that every element of P is an irre-
ducible element of R and for every irreducible element f ∈ R, there is exactly one
g ∈ P such that (f ) = (g).

A Cohen-Kaplansky domain is an atomic domain with finitely many irreducibles.


In §15.11 we showed that a Cohen-Kaplansky domain R that is not a field is a
semilocal one-dimensional Noetherian domain. Let R be the integral closure of R
in its fraction field K. By Krull-Akizuki the ring R is a Dedekind domain, and by
Corollary 18.8, the ring R has only finitely many prime ideals, so R is a semilocal
PID.
Exercise 22.6. Let R be a PID. Show: R is a Cohen-Kaplansky domain iff
MaxSpec R is finite, in which case R has # MaxSpec R irreducibles.
However, in Example 15.51 we saw that a semilocal one-dimensional Noetherian
domain may or may not be Cohen-Kaplansky. And now our story continues.
Example 22.6. Let R be a primefree Cohen-Kaplansky domain which is not
a field. We claim that any maximal ideal m of R has at least three irreducibles.
By Lemma 15.49, m is generated by its irreducibles; if it were generated by one
irreducible we would have a prime element, so m must contain nonassociate irre-
ducibles f1 ,f2 . Then f1 + f2 is a nonzero element of m so has an irreducible factor
f3 ∈ m, which cannot be associate to either f1 or f2 .
Suppose now that (R, m) is a local Cohen-Kaplansky domain with exactly three
irreducibles f1 , f2 , f3 . Then f2 f3 + f1 ∈ m• and is not divisible by f2 or f3 , so
it follows that there is α ∈ R× and a ∈ Z+ with f2 f3 = αf1a . Similarly we find
β, γ ∈ R× and b, c ∈ Z+ with f3 f1 = βf2b and f1 f2 = γf3c . Multiplying these
relations we get
f12 f22 f32 ∼ f1a f2b f3c
from which it follows (without using uniqueness of factorization!) that a = b = c =
2. Thus the multiplicative structure of such an R is completely determined.
Lemma 22.7. Let R be a semilocal ring, with MaxSpec R = {p1 , . . . , pm }.
a) We have that R is Noetherian iff Rpi is Noetherian for all 1 ≤ i ≤ m.
b) We have dim R = maxj dim Rpj .
1Suggestion: show R is residually finite, and use properties of the norm function ||I|| = #R/I.
2. COHEN-KAPLANSKY DOMAINS 355

Proof. a) If R is Noetherian, then so are all of its localizations, hence in


particular Rpj for all 1 ≤ j ≤ m. Suppose R is not Noetherian, and let {In }∞ n=1 be
an infinite strictly ascending chain of ideals. To each In we may associate the subset
Vn of MaxSpec R of maximal ideals in which In is contained. Since In is proper,
each Vn is nonempty, Since In+1 ⊃ In , we have Vn+1 ⊂ Vn . Since MaxSpec R is
finite, it follows that there is some j such that pj ∈ Vn for all n, and thus {In } is
still an infinite strictly ascending chain in Rpj , which is thus not Noetherian.
b) Every chain of prime ideals of R can be extended to a chain of prime ideals
terminating at pj for some j, which is a chain of prime ideals in Rpj . The result
follows. 

In any one-dimensional domain, a nonzero ideal I is primary iff r(I) is prime iff r(I)
is maximal iff I is contained in a unique maximal ideal of R. Indeed, by Proposition
10.15, if I is primary, then r(I) ⊃ I ) (0) is prime, so r(I) = m is maximal and
thus m is the only maximal ideal containing I, whereas if I is contained in exactly
one maximal ideal m, then since nonzero primes are maximal, we have r(I) = m
and I is primary by Proposition 10.14.
Now suppose R is a semi-local one-dimensional domain, with MaxSpec R =
{p1 , . . . , pm }. For a nonzero ideal I of R, let ιj : R → Rpj be the localization map,
and put
Ij = ι∗j (ιj )∗ I.
If I is not contained in pj then (ιj )∗ I = Rpj so Ij = R. Suppose that I ⊂ pj . Then
Ii ⊂ ι∗j (ιj )∗ pj = pj . Moreover, if x ∈ pj , then since r((ιj )∗ I) = pj + Rpj , there is
n ∈ Z+ such that ιj (xn ) = ιj (x)n ∈ (ιj )∗ I, so xn ∈ Ij . It follows that r(Ij ) ⊃ pj .
Thus r(Ij ) = pj and Ij is pj -primary.
Consider the family of ideals {Ij }m j=1 . Since the radicals are pairwise comaxi-
mal, so are the Ij ’s, and thus
m
Y m
\
Ij = Ij .
j=1 j=1
Tm
Moreover, consider the injection ϕ : I ,→ j=1 Ij . For all 1 ≤ j ≤ m, upon
tensoring to Rpj , both sides are pj +Rpj if I ⊂ pj and Rpj otherwise. By Proposition
??, ϕ is an isomorphism, and thus we get a primary decomposition
m
Y m
\
(45) I= Ii = Ii
i=1 i=1

which is unique by Theorem 10.28.

Suppose now that I = (a) is principal. Then each Ii is invertible, hence princi-
pal since R is semilocal (Corollary ??), say Ii = (ai ). Thus there is u ∈ R× such
that
a = ua1 · · · an , ai ∈ Ii .
Each ai is either a unit in Rpi or pi -primary, and for all i we have
(a) + pi = (ai ) + pi .
So if a is irreducible in R, then there is a unique
S j such that aj is irreducible and ai
is a unit for all i 6= j and thus that a ∈ pj \ i6=j pi . More generally, factorization
356 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

in R can be completely understood in terms of factorization in Rpi for all i. The


precise statement is as follows.
Theorem 22.8. Let R be a one-dimensional semi-local domain, with fraction
field K, and let MaxSpec R = {p1 , . . . , pm }. a) The map
Φ : x ∈ K × /R× 7→ (xRp×1 , . . . , xRp×m )
induces an order-preserving group isomorphism
m

M
G(R) → G(Rpi ).
i=1

b) There is an induced order-preserving monoid isomorphism


n

M
G+ (R) → G+ (Rpi ).
i=1

Proof. T a) The map Φ is an order-preserving group homomorphism. Moreover


m
its kernel is i=1 (R \ pj ) = R× , so it is injective. For surjectivity, it is enough to
show that for all 1 ≤ j ≤ m and all xj ∈ Rp•j there is x ∈ R• such that Φ(x)
has j-component xj and all other
Lm components 0 (i.e., lying in Rp×i ): the subgroup
y
generated by such elements is i=1 G(Rpj ). Write xj = s with s ∈ R \ pj . Perform
the primary decomposition of the principal ideal (y) as above: for all 1 ≤ i ≤ m,
((y) + Rpi ) ∩ R = (yi )
and take x = yj . Then the jth component of Φ(x) is xj Rp•j and every other com-
ponent is a unit in Rp×i .
b) Any isomorphism of partially ordered commutative groups induces an isomor-
phism on positive
Ln cones.L And for any partially ordered commutative groups G1 , . . . , Gm
n
we have ( i=1 Gi )+ = i=1 G+ i . Thus the result follows from part a). 
Corollary 22.9. Let R be a one-dimensional semilocal domain, with MaxSpec R =
{p1 , . . . , pm }.
a) Each irreducible element of R belongs to exactly one maximal ideal of R.
b) For each 1 ≤ j ≤ m, the localization map R → Rpj induces a bijection from the
atoms contained in pj to the atoms of Rpj .
∼ Lm
Proof. The ordered monoid isomorphism Φ : G+ (R) → i=1 G+ (Rpj of The-
Lm
orem X.X induces a bijection from atoms of G+ (R) to atoms of i=1 G+ (Rpj . The
atoms of any direct sum of partially ordered commutative groups consist precisely
of the tuples that are the identity in all coordinates except for one and an atom in
the remaining coordinate. The result follows. 
Corollary 22.10. Let R be a semilocal domain, with MaxSpec R = {p1 , . . . , pm }.
Then R is a Cohen-Kaplansky domain iff Rpj is a Cohen-Kaplansky domain for all
1 ≤ j ≤ m.
Proof. If R is a Cohen-Kaplansky domain, then R is Noetherian of dimension
one, hence so is each Rpj . By Corollary 22.9, since R has finitely many atoms, so
does each Rpj . So each Rpj is Cohen-Kaplansky.
Suppose each Rpj is Cohen-Kaplansky. Then each Rpj is Noetherian of dimen-
sion one, so by Lemma 22.7 R is Noetherian of dimension on, and in particular
2. COHEN-KAPLANSKY DOMAINS 357

atomic. Since each Rpj has finitely many atoms, by Corollary 22.9 R has finitely
many atoms. So R is Cohen-Kaplansky. 
In any local domain (R, m) an element of f ∈ m \ m2 must be irreducible. Such an
element induces a nonzero element of the Zariski cotangent space m/m2 , which
is a vector space over the residue field k = R/m; let us write f for the element
f + m ∈ m/m2 . If f, g ∈ m \ m2 are associate elements, then f and g generate the
same one-dimensional k-subspace of m/m2 . (But the converse need not be true.)
We have shown the following fact.
Proposition 22.11. Let (R, m) be a Noetherian local domain. Then R has at
least as many nonassociate irreducibles as elements of P(m/m2 ), the projectivized
Zariski cotangent space.
Exercise 22.7. Let (R, m) be a Noetherian local domain with residue field
k = R/m. Suppose R is not a DVR. Show: R has at least #k + 1 irreducibles. In
particular, if R is a Cohen-Kaplansky domain then k is finite.
Let (R, m) be a primefree local Cohen-Kaplansky domain which is not a DVR,
with k = R/m. Thus k ∼ = Fq Thus d = dimk m/m2 > 1. By X.X R has at least
d
−1
#P d−1 (Fq ) = qq−1 -irreducibles. Our next order of business is to understand when
we have equality.

Let R be a domain. An element x ∈ R is universal if it is divisible by every


irreducible element of R. The set of all universal elements of R forms an ideal,
which we will denote by u. In fact this ideal is precisely the intersection of the
principal ideals generated by all irreducible elements. We call an ideal I universal
if it is contained in u.
Exercise 22.8. a) Let (R, m) be a Noetherian local domain which is not a field.
Show: u = m iff R is a DVR.
b) Let R be a Cohen-Kaplansky domain. Show: the ideal u of all universal elements
is nonzero.
c) Let R be a domain in which the ideal u of universal elements is not zero. Must
R be a Cohen-Kaplansky domain?
Theorem 22.12. Let (R, m) be a primefree local Cohen-Kaplansky domain with
k = R/m ∼= Fq and d = dimk m/m2 .
d
−1
a) R has exactly qq−1 irreducibles iff m2 is universal.
q d+1 −1
b) If m2 contains an irreducible element then R has at least q−1 irreducibles.
d
−1
Proof. a) We have more than qq−1 irreducibles iff either some element of m2
is irreducible or there are nonassociate irreducibles f, g ∈ m and u ∈ R× such that
f − ug ∈ m2 . If m2 contains an irreducible element then m2 is not universal. If
there are nonassociate irreducibles f, g ∈ m and u ∈ R× such that f − ug ∈ m2 ,
then f − ug cannot be divisible by either f or g so m2 is not universal. Conversely,
suppose m2 is not universal and that every irreducible lies in m (otherwise we know
d
−1
we have more than qq−1 irreducibles). Since m2 is generated by the set of products
of two irreducibles, there are (not necessarily distinct) irreducibles f1 , f2 , f3 such
that f3 - f1 f2 . Since f1 f2 + f3 ∈ m, it has an irreducible divisor f4 , which cannot be
358 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

associate to f3 . Let x ∈ R be such that f4 x = f1 f2 + f3 . We must have x ∈ R× for


otherwise f3 ∈ m2 , and we find that f3 and f4 are nonassociate irreducibles which
generate the same one-dimensional subspace of m/m2 .
b) COMPLETE ME! 
2.1. Groups of divisibility.

Let R be a domain with fraction field K and integral closure R. The group
of divisibility G(R) = K × /R× , partially ordered under divisibility.
Exercise 22.9. Show: R is a Cohen-Kaplansky domain iff the semigroup
G(R)+ = R• /R× is finitely generated.
We have a short exact sequence
×
(46) R /R× → G(R) → G(R) → 0.
Since R is integrally closed, if x ∈ K and xn ∈ R for some n ∈ Z+ then x ∈
R: it follows that G(R) is torsionfree. IF R is Cohen-Kaplansky then G(R)+ is
a finitely generated semigroup, hence G(R) is a finitely generated group, hence
×
R /R× is finite and G(R) is finitely generated, and thus (46) splits (as a sequence
of commutative groups). Moreover if # MaxSpec R = # MaxSpec R = n, then
since R is a UFD with n primes, we have G(R) ∼ = Zn . Thus finally we have an
isomorphism of commutative groups
n
×
G(R) ∼
M
= R /R× ⊕ Z.
i=1

Theorem 22.13. a) For a domain R, the following are equivalent:


(i) G(R) is a finitely generated commutative group.
(ii) R is semilocal and G(Rm ) is finitely generated for all m ∈ MaxSpec R.
b) Let R be a one-dimensional semilocal domain with maximal ideals m1 , . . . , mn .
The natural map
Mn
G(R) → G(Rmi )
i=1
is an order isomorphism. It follows that
n
G(R)+ ∼
M
= G(Rmi )+ .
i=1

Corollary 22.14. For a domain R, the following are equivalent:


(i) R is Cohen-Kaplansky and # MaxSpec R = n.
(ii) G(R) ∼= T ⊕ Zn with T finite.
(iii) R is Noetherian, G(R) is finitely generated and # Max R = #M axR.
2.2. Composites.

An extension of domains R ⊂ S is called radical if for all s ∈ S there is n =


n(s) ∈ Z+ with sn(s) ∈ R. In particular radical extensions are integral.
Proposition 22.15. If ι : R ⊂ S is radical, then ι∗ : Spec S → Spec R is a
homeomorphism. It follows that ι∗ : MaxSpec S → MaxSpec R is a bijection.
3. RINGS OF FINITE RANK 359

3. Rings of Finite Rank


For a module M over a ring R, denote by µ(M ) the minimal cardinality of a set
of R-module generators for M . This is a cardinal number, bounded above by #M .
We define µ∗ (M ) to be the supremum of µ(N ) as N ranges over all R-submodules of
M . Finally, define the rank of R, denoted rk(R), as µ∗ (R), i.e., as the supremum
of µ(I) as I ranges over all ideals of R. We say R has finite rank if rk(R) < ℵ0 .
Exercise 22.10. a) Show: R is Noetherian iff µ(I) < ℵ0 for all ideals I of R.
b) Show: if rk(R) < ℵ0 , then R is Noetherian.
c) Must a Noetherian ring have finite rank?
(This is a difficult question to answer from scratch. We will soon give an an-
swer...but using results from dimension theory that we do not develop in this text.)
Example 22.16. Let R be a ring.
a) We have rk(R) = 0 iff R is the zero ring.
b) We have rk(R) = 1 iff R is a nonzero principal ring.
c) Let R be a Dedekind domain and not a PID. Theorem 20.12 implies rk(R) = 2.
Exercise 22.11. Let ι : R → S be a homomorphism of rings.
a) Let M be an R-module, and let ι∗ (M ) be the S-module S ⊗R M . Show:
µ(ι∗ (M )) ≤ µ(M ).
b) Suppose every ideal of S is extended from R: i.e., every ideal J of S is of the
form ι∗ (I) for some ideal I of R. Show:
rk(S) ≤ rk(R).
Deduce that this holds if ι is a quotient map or a localization map.
Exercise 22.12. a) Let R = R1 × R2 be a direct product of rings. Let M be
an R-module. Show: there is an R1 -module M1 and an R2 -module M2 such that
M is isomorphic to M1 × M2 as an R1 × R2 -module. Deduce:
µ(M ) = max µ(M1 ), µ(M2 ).
Qn
b) Show: if R = i=1 Ri is a finite direct product of rings, then
rk(R) = max rk(Ri ).
1≤i≤n

Exercise 22.13. Show: rk(Z[ −3]) = 2.
Exercise 22.14. Let R be an Artinian ring.
a) Show rk(R) ≤ `(R). Deduce: R has finite rank.
b) Suppose R is not a field. Show: rk(R) ≤ `(R) − 1.
3.1. Cohen.
Exercise 22.15. Let Fq be a finite field, and let R = Fq [t1 , . . . , tn ]/ht1 , . . . , tn i2 .
a) Show: R is a local ring with maximal ideal m = ht1 , . . . , tn i + ht1 , . . . , tn , i2 .
b) Show: R/m = Fq , #m = q n and #R = q n+1 .
c) Show: µ(m) = n and thus rk(R) ≥ n.
d) What is rk(R)?
A ring R is residually Artinian if R/I is Artinian for all nonzero ideals I of R.
360 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

Exercise 22.16. Show:


a) An Artinian ring is residually Artinian.
b) A residually Artinian ring that is not a domain is Artinian.
c) If R is residually Artinian, then dim R ≤ 1.
d) A residually Artinian ring is Noetherian.
Proposition 22.17. A domain R is residually Artinian iff it is Noetherian
and dim R ≤ 1.
Proof. It follows from the previous exericse that a residually Artinian domain
is Noetherian of dimension at most one. Conversely, suppose R is a Noetherian
domain of dimension at most 1. If dim R = 0 then R is a field, hence is certainly
residually
Tn Artinian, so suppose dim R = 1 and let I be a nonzero ideal in R. Let
I = i=1 qi be a primary decomposition of I, and let pi = r(qi ). Then as in §10.8,
each pi is maximal, so by the Chinese Remainder Theorem we have
n
R/I ∼
Y
= R/qi .
i=1

Moreover each R/qi is Noetherian with unique prime ideal pi , so is a local Artinian
ring. It follows that R/I is Artinian. 
Exercise 22.17. (Cohen) Let R be a residually Artinian domain with fraction
field F , let K/F be a finite degree field extension, and let S be the integral closure
of R in K. Show: S is residually Artinian.
We warn the reader that the proof of the following result uses three different aspects
of commutative algebra that are not covered in these notes.
Theorem 22.18. (Cohen [Co50])
a) For a local domain (R, m), the following are equivalent:
(i) R has finite rank.
(ii) R is Noetherian of dimension at most 1.
b) If R has finite rank, then dim R ≤ 1.
Proof. a) ¬ (ii) =⇒ ¬ (i): If R is not Noetherian, then certainly rk R ≥ ℵ0 ,
so we may assume that R is Noetherian and of Krull dimension at least 2 and show
that R does not have finite rank. By the Generalized Principal Ideal Theorem,
we have 2 ≤ ht m < ℵ0 . By the theory of Hilbert polynomials – not covered here
(sorry!) but see [AM, Ch. 11] – there is N ∈ Z+ and a degree d − 1 polynomial
P (t) ∈ Q[t] such that for all n ≥ N we have
dimR/m mn /mn+1 = P (n).
It follows that
lim µ(mn ) = lim P (n) = ∞
n→∞ n→∞
and thus rk(R) = ℵ0 .
(ii) =⇒ (i): Let (R, m) be a Noetherian local domain of dimension at most 1. By
Proposition 22.17, the ring R is residually Artinian. Let R be the m-adic completion
of R. By Cohen’s results on the structure of complete local rings [Co46], there is a
subring R0 of R such that R0 is a complete discrete valuation ring and R is finitely
generated free R-module, say of rank d. By the standard PID structure theory,
every ideal I of R is then an R0 -submodule of the rank d R0 -module R, hence is
3. RINGS OF FINITE RANK 361

itself free of rank at most d as an R0 -module and thus a fortiori can be generated
as an ideal by at most d elements: that is, rk R ≤ d.
Now let I be a nonzero ideal of R. Then IR can be generated by d elements
a1 , . . . , ad . Since m is the only proper prime ideal of R, mI is m-primary, hence
mI ⊃ mN for some N ∈ Z+ . Let a1 , . . . , ad ∈ R be such that
ai ≡ ai (mod mIR)∀1 ≤ i ≤ d.
We have ai ∈ IR; since IR ∩ R = I, we have ai ∈ I. Let J := ha1 , . . . , ad iR , so
J ⊂ I and µ(J) = d. Thus it suffices to show that J = I. We have
IR ⊂ JR + ImR,
so
I ⊂ (JR + ImR) ∩ R = J + Im,
so J = I by Corollary 3.42.
b) Suppose R has finite rank. Then R is Noetherian. Seeking a contradiction, we
assume dim R ≥ 2, i.e., that there are prime ideals p0 ( p1 ( p2 of R. Let S be
the localization of R/p0 at p2 , so S is Noetherian local and dim S ≥ 2. On the
one hand, by part a) S does not have finite rank. On the other hand, since S is
obtained from the finite rank ring S by a quotient map followed by a localization,
S has finite rank: contradiction. 

3.2. Gilmer.
Proposition 22.19. (Gilmer) Let M be an R-module, and let N be an R-
submodule. Let r, s be cardinal numbers.
a) If µ∗ (M ) ≤ r, then µ∗ (N ) ≤ r and µ∗ (M/N ) ≤ r.
b) If µ∗ (N ) ≤
Lrn and µ∗ (M/N ) ≤ s, then µ∗ (M ) ≤ r + s.
c) If M = i=1 Mi is a finite direct sum of R-modules, then µ∗ (M ) < ℵ0 iff
µ∗ (Mi ) < ℵ0 for all i.
Proof. Let q : M → M/N be the natural map.
a) Suppose µ∗ (M ) ≤ r, so µ(N 0 ) ≤ r for all submodules N 0 of M . Since every
submodule of N is also a submodule of M , we deduce µ∗ (N ) ≤ r. Similarly, if
P is a submodule of M/N , then P = q −1 (P )/N , and since µ(q −1 (P )) ≤ r, also
µ(P ) ≤ r and thus µ∗ (M/N ) ≤ r.
b) Suppose µ∗ (N ) ≤ r and µ∗ (M/N ) ≤ s, and let N 0 be a submodule of N . Then
µ(q(N 0 )) ≤ s, so let S ⊂ N 0 of cardinality at most s mapping to a set of generators
of q(N 0 ). Let R be a set of generators of N ∩ N 0 of cardinality at most R. Then
R ∪ S generates N and has cardinality at most r + s.
c) This follows from parts a) and b). 

Proposition 22.20. (Gilmer) Let M be a finitely generated R-module. Then


µ∗ (M ) ≤ µ(M ) rk(R).
Proof. We may assume M 6= 0. We proceed by induction on k = µ(M ) ∈ Z+ .
Base Case (k = 1): If M is monogenic, then M ∼
= R/ ann(M ), so
µ∗ (M ) ≤ µ∗ (R) = rk(R).
Induction Step: suppose k ≥ 2 and the result holds for all modules M with µ(M ) ≤
k − 1, let M be a module with µ(M ) = k and put M = hx1 , . . . , xk i. Put N =
362 22. ONE DIMENSIONAL NOETHERIAN DOMAINS

hx1 , . . . , xk−1 i, so µ(N ) ≤ k − 1 and µ(M/N ) ≤ 1. By induction we have µ∗ (N ) ≤


µ(N ) rk(R) and µ∗ (M/N ) ≤ rk(R), so
µ∗ (M ) ≤ (µ(N ) + 1) rk R ≤ µ(M ) rk(R). 
Theorem 22.21. (Gilmer) For R Noetherian, the following are equivalent:
(i) R has finite rank.
(iI) For all minimal primes p ∈ Spec R, the ring R/p has finite rank.
Proof. (i) =⇒ (ii) is clear from Exercise 22.11.
(ii) =⇒ (i): Step 1: We show that for each minimal p ∈ Spec R and n ∈ Z+ ,
the ring R/pn has finite rank. We proceed by induction on n, and the base case is
our hypothesis that R/p has finite rank. Let n ∈ Z+ and suppose rk R/pn is finite.
Notice that since R → R/pn is surjective, the rank of f R/pn as a ring is the same
as its rank µ∗ (R/pn ) as an R-module. Consider the exact sequence of R-modules
0 → pn /pn+1 → R/pn+1 → R/pn → 0.
By Proposition 22.19 we have
µ∗ (R/pn+1 ) ≤ µ∗ (pn /pn+1 ) + µ∗ (R/pn ).
The rank of pn /pn+1 as an R-module is equal to its rank as an R/p-module. Thus
by X.X we have
µ∗ (pn /pn+1 ) ≤ rk(R/p) rk(pn /pn+1 ) < ℵ0
since R is Noetherian. Thus overal we have shown rk(R/pn+1 ) = µ∗ (R/pn+1 ) < ℵ0 .
Step 2: Because R is Noetherian, it has finitely many
Tr minimal primes p1 , . . . , pr ,
and by Exercise 10.13 there is N ∈ Z+ such that i=1 pN i = (0). Thus we have an
injection of R-modules
Mr
R ,→ R/pNi ,
i=1
and it follows from Step 1 and Proposition 22.19 that rk(R) < ℵ0 . 
3.3. Further results.

In studying rings of finite rank, it seems reasonable to assume that R is Noe-


therian – for otherwise rk R ≥ ℵ0 – and nonprincipal – for otherwise rk R ≤ 1. (Of
course this presumes that we can recognize Noetherian and principal rings when
we encounter them. The former seems reasonable in view of our large number of
characterizations of Noetherian rings. The latter is in practice quite dubious – e.g.
it is completely unclear when the integral closure ZK of Z in a number field K is a
principal ring – but from a general perspective it seems more reasonable.)
Theorem 22.22. Let R be a Noetherian, nonprincipal ring.
a) If rk R < ℵ0 , then dim R ≤ 1.
b) We have
(47) rk R ≤ sup (rk Rm + 1)
m∈MaxSpec R

and thus [Gi72, Thm. 3] rk R < ℵ0 iff supm∈MaxSpec R rk Rm < ∞.


c) Suppose R is moreover reduced. Then rk R can be computed locally:
(48) rk R = sup rk Rm .
m∈MaxSpec R
3. RINGS OF FINITE RANK 363

Proof. a) We argue by contraposition: suppose dim R ≥ 2. Let p be a minimal


prime ideal of R. Then R/p is a domain of dimension at least 2, so by Theorem
22.18 we have rk R/p = ℵ0 and thus by Exercise 22.11 we have rk R = ℵ0 .
b) Let I be an ideal of R. By the Forster-Swan Theorem, we have
µ(I) ≤ sup (µ(IRp ) + dim JSpec R/p) ≤ sup (µ(IRp ) + dim R/p).
p∈JSpec R p∈Spec R

We may assume that dim R ≤ 1: otherwise dim Rm ≥ 2 for some m ∈ MaxSpec R,


so by part a) we have rk Rm = ℵ0 which then implies rk R = ℵ0 and thus (48)
holds. Thus p ∈ Spec R is either maximal or minimal. If p = m is maximal we have
µ(IRm ) + dim R/m = µ(IRm ) ≤ rk Rm . If p is minimal, let m be a maximal ideal
containing p, so
µ(IRp ) + dim R/p ≤ µ(IRm ) + 1 ≤ rk Rm + 1,
which establishes (47).
c) If R is reduced, then Rp is a reduced Artinian local ring, i.e., a field, so
µ(IRp ) + dim R/p ≤ 2,
but since R is nonprincipal, we have rk R ≥ 2. This shows that
rk R ≤ sup rk Rm ,
m∈MaxSpec R

but again we certainly have rk Rm ≤ rk R for all m, so (48) follows. 


CHAPTER 23

Structure of overrings

1. Introducing overrings
Let R be a domain with fraction field K. Let ΣR be the set of height one prime
ideals of R. By an overring of R we mean a subring of K containing R, i.e., a
ring T with R ⊂ T ⊂ K. (We allow equality.) This is standard terminology among
commutative algebraists, but we warn that someone who has not heard it before
will probably guess incorrectly at its meaning: one might well think that “T is an
overring of R” would simply mean that “R is a subring of T ”.

2. Overrings of Dedekind domains


In this section we will study overrings of a Dedekind domain R. In particular, we
will answer the following questions.
Question 5. Let R be a Dedekind domain.
a) Can we (in some sense) classify the overrings of R?
b) Under what conditions is every overring of R a localization?
c) Let T be an overring of R. What is the relationship between Pic T and Pic R?
(In point of fact we can ask these questions for any domain R. But for a Dedekind
domain we will obtain definitive, useful answers.)

As a warmup, suppose R is a PID. In this case every overring is indeed a localiza-


tion: to see this it is enough to show that for all coprime x, y ∈ R• , y1 ∈ R[ xy ]. But
since x and y are coprime  in the PID R, there are a, b ∈ R such that ax + by = 1,
and then y1 = ax+by
y = a x x
y + b ∈ R[ y ]. It follows that every overring of a PID is
obtained by localizing at a multiplicative subset S ⊂ R• . Further, by uniqueness
of factorization the saturated multiplicatively closed subsets of R• are in bijection
with subsets of ΣR : in other words, an overring is entirely determined by the set
of prime elements we invert, and inverting different sets of prime elements leads to
distinct overrings. Further, since a localization of a PID is a again a PID, in this
case we have Pic T = 0 for all overrings.

Some of the above analysis generalizes to arbitrary Dedekind domains: we will


show that for any Dedekind domain R the overrings of R correspond bijectively to
subsets of ΣR . More precisely, for any subset W ⊂ ΣR we define
\
RW = Rp
p∈W

365
366 23. STRUCTURE OF OVERRINGS

and also \
RW = Rp .
p∈ΣR \W

(Note that RW = RΣR \W . So it is logically unnecessary to consider both RW and


RW , but it will be notationally convenient to do so.) When W = {p} consists of
a single element, we we write Rp for R{p} . Let Φ : ΣR → Pic R be the ideal class
map, i.e., p 7→ [p]. Then we will show that every overring of a Dedekind domain R
is of the form RW for a unique subset W ⊂ ΣR and that Pic RW ∼ = Pic R/hΦ(W ).
On the other hand, it is not generally true that an arbitrary overring of R arsies
as a localization. However, O. Goldman gave a beautiful analysis of the situation:
it turns out that a Noetherian domain R has the property that every overring is a
localization iff R is a Dedekind domain and Pic R is a torsion group.

2.1. Structure of overrings. Let R be a Dedekind domain with fraction


field K.
Lemma 23.1. a) For all p ∈ ΣR , there exists fp ∈ Rp \ R.
b) The mapping from subsets of ΣR to overrings of R given by W 7→ RW is injective.
Proof. a) Choose x ∈ K \ Rp , and let S be the finite set of maximal ideals
q distinct from p such that ordq (x) <Q0. For each q ∈ S, let yq ∈ q \ p. Let
N = maxq∈S − ordq (x), and put fp = ( q∈S yq )N x.
b) Suppose W1 and W2 are distinct subsets of ΣR . After relabelling if necessary,
we may assume that there exists p ∈ W2 \ W1 . By part a), there exists fp ∈ Rp \ R
and thus fp ∈ RW1 \ RW2 . 
Proposition 23.2. Let R be a Dedekind domain with fraction field K, and let
T be an overring of R. Write ι : R ,→ T for the inclusion map.
a) For every P ∈ ΣT , T = RP∩R .
b) T is itself a Dedekind domain.
c) ι∗ : ΣT ,→ ΣR is an injection.
d) For all P ∈ ΣT , ι∗ ι∗ P = P.
e) ι∗ identifies ΣT with the subset of p ∈ ΣR such that pT ( T .
Proof. a) Put p = P ∩ R. There exist x, y ∈ R• such that xy ∈ P. Then
0 6= x = y( xy ) ∈ p, so p is a nonzero prime ideal of R. Thus TP contains the DVR
Rp and is properly contained in its fraction field, so TP = Rp .
b) By the Krull-Akizuki Theorem, T is a Noetherian domain of Krull dimension
one. By part a), the localization of T at every prime is a DVR. So T is integrally
closed and is thus a Dedekind domain.
c) For distinct P1 , P2 ∈ ΣT , the localizations TP1 , TP2 are distinct DVRs. But by
part a), putting pi = ι∗ (Pi ), we have Rpi = TPi for i = 1, 2, so p1 6= p2 .
d) Suppose p = ι∗ (P). Then
(ι∗ ι ∗ P)TP = pTP = PTP .
By part c), ι ∗ ι∗ P is not divisible by any prime other than P, so ι∗ p = P.
e) This follows immediately and is stated separately for later use. 
Theorem 23.3. For any overring T of the Dedekind domain R, we have
T = Rι∗ (ΣT ) .
2. OVERRINGS OF DEDEKIND DOMAINS 367

The proof (that I know) of this theorem requires the study overrings of more general
Pr̈ufer domains, to which we turn in the next section.

2.2. When overrings are localizations.


Lemma 23.4. Let R be an integrally closed domain with fraction field K, and
let T be an overring of R.
a) The relative unit group T × /R× is torsionfree.
b) Suppose that R is a Dedekind domain, p ∈ ΣR and T = Rp . the following are
equivalent:
(i) T × /R× ∼= Z.
(ii) T × ) R× .
(iii) [p] ∈ Pic(R)[tors].
(iv) There exists x ∈ R which is contained in p and in no maximal ideal q 6= p.
Proof. a) Since R is integrally closed, all finite order elements of K × (i.e.,
roots of unity in K) lie in R and a fortiori in T : R× [tors] = T × [tors]. On the other
hand, let x ∈ T × be of infinite order such that xn ∈ R× for some n ∈ Z+ . Again
integral closure of R implies x ∈ R, and then xn ∈ R× =⇒ x ∈ R× .
b) (i) =⇒ (ii) is clear.
(ii) =⇒ (iii): Let x ∈ K × . Then x ∈ T × iff Rx = pa for some a ∈ Z, and x ∈ R×
iff a = 0. Therefore (ii) holds iff some power of p is principal, which is to say that
the class of p ∈ Pic R is torsion.
(iii) =⇒ (i): Let a be the least positive integer such that pa is principal. Thus
pa = xR with x uniquely determined modulo R× . It follows that T × is generated
by R× and x, so T × /R× is a nontrivial cyclic group. By part a) it is also torsionfree
so T × /R× ∼= Z.
(iii) =⇒ (iv): If pa = xR, then x lies in p but in no other maximal ideal q.
(iv) =⇒ (iii): If a = vp (x), then a > 0 and xR = pa . 
Remark: Part (iv) of Lemma 23.4 was added following an observation of H. Knaf.
Theorem 23.5. (Goldman [Gol64]) For a Dedekind domain R, the following
are equivalent:
(i) Pic R is a torsion group.
(ii) Every overring of R is a localization.
Proof. (i) =⇒ (ii): Let p ∈ ΣR . By Lemma 23.1 Rp is a proper overring
of R, so by assumption Rp is a localization of R and thus has a strictly larger
unit group. By Lemma 23.4 this implies that [p] ∈ Pic(R)[tors]. Since Pic(R) is
generated by the classes of the nonzero prime ideals, it follows that Pic R is torsion.
(ii) =⇒ (i): Let T be an overring of R, and put S = R∩T × . We want to show that
T = S −1 R. That S −1 R ⊂ T is clear. Conversely, let x ∈ T , and write xR = ab−1
with a, b coprime integral ideals of R: a + b = R. Thus aT + bT = T whereas
aT = xbT ⊂ bT , so bT = T and hence also bn T = T for all n ∈ Z+ . Since Pic R is
torsion, there exists n ∈ Z+ with bn = bR. It follows that bT = T and thus b ∈ S.
Now xb = a ⊂ R, so xb ∈ R. Thus x ∈ S −1 R, and we conclude T ⊂ S −1 R. 
Corollary 23.6. Suppose that W is a finite subset of ΣR and that every
p ∈ W has finite order in Pic R. Then there is a ∈ R• such that RW = R[ a1 ].
Exercise 23.1. Prove Corollary 23.6.
368 23. STRUCTURE OF OVERRINGS

Exercise 23.2. Let T = RW be an overring of R such that T = R[ a1 ] for some


a ∈ R• .
s) Show: W is finite.
b) Must it be the case that every p ∈ W has finite order in Pic R?
2.3. The Picard group of an overring.
Theorem 23.7. Let R be a Dedekind domain, W ⊂ ΣR , let RW = p∈ΣR \W Rp ,
T
L
and let FracW R = p∈W Z denote the subgroup of fractional R-ideals supported
on W . There is a short exact sequence
v ι
1 → R× → (RW )× −→ FracW R → Pic R −→

Pic RW → 1.
Proof. The map v : (RW )× → FracW R is obtained by restricting the canon-
ical map K × → Frac R to (RW )× : the fractional ideals so obtained have p-adic
valuation 0 for all p ∈ MaxSpec RW = MaxSpec R \ W : thus the image lands in
FracW R.
It is easy to see most of the exactness claims: certainly R× → (RW )× is injective;
further, for x ∈ (RW )× , v(x) = 0 iff vp (x) = 0 for all p ∈ W ∪ MaxSpec RW =
MaxSpec R iff x ∈ R× . If I ∈ FracW R, then I is principal iff it has a genera-
tor x ∈ K × with vp (x) = 0∀p ∈ MaxSpec R \ W = MaxSpec RW iff I = (x) for
x ∈ (RW )× . Exactness at Pic R: Let [I] ∈ Pic R be such that ι∗ ([I]) = 1: thus there
is x ∈ K × with IRW = xRW . Then [I] = [x−1 I] and x−1 I ∈ FracW R. Conversely,
if I ∈ FracW R, then IRW = RW . Finally, by Proposition 23.2 ι∗ ◦ ι∗ = 1Σ(S) , so
every prime ideal of RW is of the form ι∗ (p) for a prime ideal of R. This certainly
implies that ι∗ : ι∗ : Pic R → Pic RW is surjective. 
Exercise 23.3. We maintain the setup of Theorem 23.7.
a) Use Theorem 23.7 to give a new proof of Lemma 23.4.
a) Show that the relative unit group (RW )× /R× is free abelian. (This strengthens
Lemma 23.4 when R is a Dedekind domain.)
b) Suppose Pic R is torsion. Show:
(RW )× ∼
M
= R× ⊕ Z.
p∈W

c) Deduce the Hasse-Chevalley S-Unit Theorem from the Dirichlet Unit Theorem.
d) Suppose that K is a number field. Show that K × is isomorphic to the product
of a finite cyclic group with a free commutative group of countable rank.
2.4. Repleteness in Dedekind domains.

Let R be a Dedekind domain, and consider the map Φ : ΣR → Pic R given by


p 7→ [p]. We say that R is replete if Φ is surjective, i.e., if every element of Pic R
is of the form [p] for some prime ideal p.

Example: Let R be an S-integer ring in a global field. It follows immediately


from the Chebotarev Density Theorem that R is replete.

For our coming applications it is useful to consider a variant: we say that a Dedekind
domain R is weakly replete if for every subgroup H ⊂ Pic R, there is a subset
WH ⊂ ΣR such that hΦ(WH )i = H. The point of this condition is that it allows a
complete classification of the Picard groups of overrings of R. Indeed:
3. ELASTICITY IN REPLETE DEDEKIND DOMAINS 369

Proposition 23.8. Let R be a weakly replete Dedekind domain. Then for any
subgroup H of Pic R, there is an overring T of R such that Pic T ∼
= (Pic R)/H.
Proof. By definition of weakly replete, there exists a subset W ⊂ ΣR such
that hΦ(W )i = H. By Theorem 23.7, Pic RW ∼ = Pic R/hΦ(W )i ∼ = (Pic R)/H. 

Proposition 23.9. Let R be a Dedekind domain and RW an overring of R.


a) If R is replete, so is RW .
b) If R is weakly replete, so is RW .
Exercise 23.4. Prove Proposition 23.9.
A repletion of a Dedekind domain R is a replete Dedekind domain S together

with an injective ring homomorphism ι : R ,→ S, such that ι∗ : Pic(R) → Pic(S).
Theorem 23.10. (Claborn) For a Dedekind domain R, let R1 denote the local-
ization of R[t] at the multiplicative set generated by all monic polynomials. Then
R1 is Dedekind and the composite map ι : R → R[t] → R1 is a repletion.
Proof. COMPLETE ME! 

Exercise 23.5. Use Proposition 23.8 and Theorem 23.10 to show: if for every
cardinal κ, there is a Dedekind domain R with Picard group a free commutative
group of rank κ, then for every commutative group G there is a Dedekind domain
R with Pic R ∼= G.

3. Elasticity in Replete Dedekind Domains


Let R be a domain and x ∈ R• \ R× . If for n ∈ Z+ there are (not necessarily
distinct) irreducible elements α1 , . . . , αn of R such that x = α1 · · · αn , we say that
x admits an irreducible factorization of length n.

A half factorial domain (or HFD) is an atomic domain in which for all x ∈
R• \ R× , any two irreducible factorizations of x have the same length.

Exercise 23.6. (Zaks) Show: Z[ −3] is a HFD which is not integrally closed.
For R an atomic domain and x ∈ R• \ R× , let L(x) be the supremum of all lengths
of irreducible factorizations of x and let `(x) be the minimum of all lengths of irre-
ducible factorizations of x. We define the elasticity of x, ρ(x), as the ratio L(x)
`(x) .
×
We also make the convention that for x ∈ R , ρ(x) = 1. Finally we define the
elasticity of R as ρ(R) = supx∈R• ρ(x).

An atomic domain is a HFD iff ρ(R) = 1. Thus ρ(R) is a quantitative measure of


how far an atomic domain is from being a HFD.

Let (G, ·) be a commutative group. A finite sequence g1 , . . . , gn ofQ elements in


G is irreducible if for all nonempty proper subsets S ⊂ {1, . . . , n}, i∈S gi 6= 1.
Lemma 23.11. Let (G, ·) be a commutative group, let x1 , . . . , xn be an irre-
Qn −1
ducible sequence in G, and let xn+1 = ( i=1 xi ) . If xn+1 6= 1, then x1 , . . . , xn , xn+1
is an irreducible sequence.
370 23. STRUCTURE OF OVERRINGS

Proof. A nontrivial proper subsequence of x1 , . . . , xn+1 with trivial prod-


uct must be of the form xi1 , . . . , xik , xn+1 for some nonempty proper subset S =
0
Q −1
{i
Q 1 , . . . , i k } of {1, . . . , n}. Put S = {1, . . . , n} \ S. Then i∈S 0 x i = 1, hence also
i∈S 0 x i = 1, contradicting the irreducibility of x 1 , . . . , x n . 
Proposition 23.12. Let R be a Dedekind domain, let x ∈ R• \ R× , and let
Yr
(x) = pi
i=1
be the factorization of x into prime ideals.
a) (Carlitz-Valenza [Ca60] [Va90]) The following are Qequivalent:
(i) For no nonempty proper subset S ⊂ {1, . . . , r} is i∈S pi is principal.
(ii) The element x is irreducible.
b) If p is a prime ideal such that pr = (x) and ps is nonprincipal for all 1 ≤ s < r,
then x is irreducible.
c) If no pi is principal, the length of any irreducible factorization of x is at most 2r .
Exercise 23.7. Prove Proposition 23.12.
For a commutative group (G, ·) the Davenport constant D(G) of G is the max-
imum length of an irreducible sequence in G, or ∞ if the lengths of irreducible
sequences in G are unbounded.
Proposition 23.13. Let R be a Dedekind domain, and let x ∈ R• be irre-
ducible. Write (x) = p1 · · · pr . Then r ≤ D(Pic R).
Proof. By Proposition 23.12a), p1 , . . . , pr is an irreducible sequence in Pic R.

Proposition 23.14.
a) If H is a subgroup of a commutative group G, then D(H) ≤ D(G).
b) If H is a quotient of a commutative group G, then D(H) ≤ D(G).
c) D(G) ≥ exp G = supx∈G #hxi.
d) If G is infinite, D(G) = ∞.
e) If G is finite, then D(G) ≤ #G.
f ) If G is finite cyclic, then D(G) = #G.
g) We have
r
M r
X
(49) D( Z/ni Z) ≥ 1 + (ni − 1).
i=1 i=1
h) We have D(G) = 1 ⇐⇒ #G = 1 and D(G) = 2 ⇐⇒ #G = 2.
Proof. a) If H is a subgroup of G, then any irreducible sequence in H is an
irreducible sequence in G.
b) If q : G → H is a surjective homomorphism and x1 , . . . , xn is irreducible in H,
then choosing any lift x̃i of xi to G yields an irreducible sequence x̃1 , . . . , x̃n .
c) If x ∈ G and n ∈ Z+ is less than or equal to the order of x, then x, x, . . . , x (n
times) is an irreducible sequence in G of length n.
d) By part c), we may assume G is infinite and of finite exponent. Then for some
prime p, G[p] is infinite, and by part a) it suffices to show that D(G[p]) = ∞.
But G[p] is an infinite-dimensional vector space over the field Fp : let {ei }∞ i=1 be an
infinite Fp -linearly independent subset of G[p]. Then for all n ∈ Z+ the sequence
3. ELASTICITY IN REPLETE DEDEKIND DOMAINS 371

e1 , . . . , en is irreducible.
e) Suppose #G = n, and let g1 , . . . , gn+1 be a sequence in G. For 1 ≤ i ≤ n, let
Pi = g1 · · · gi . By the Pigeonhole Principle there is 1 ≤ i < j ≤ n + 1 such that
Pi = Pj , and thus gi+1 · · · gj = 1.
f) Since exp Z/nZ = #Z/nZ = n, this follows from parts c) and e).
Lr Pk
g) Let G = i=1 Z/ni Z,1 and let d(G) = 1 + i=1 (ni − 1). There is an “obvious”
irreducible sequence x1 , . . . , xd(G)−1 : for 1 ≤ i ≤ k, let ei be the element of G
with ith coordinate 1 and other coordinates 0. Take e1 , . . . , e1 (n1 − 1 times),
e2 , . . . , e2 (n2 − 1 times),....,ek , . . . , ek (nk − 1 times). The sum of these elements is
Pd(G)−1
(n1 − 1, . . . , nk − 1) 6= 0, so by Lemma 23.11 taking xd(G) = − i=1 xi , we get
an irreducible sequence of length d(G).
h) Left to the reader as an easy exercise in applying some of the above parts. 
Theorem 23.15. Let R be a Dedekind domain.
a) We have ρ(R) ≤ max( D(Pic 2
R)
, 1).
b) If R is replete, then ρ(R) = max( D(Pic
2
R)
, 1).
Proof. For x ∈ R• \ R× , let P (x) be the number of prime ideals (with multi-
plicity) in the factorization of (x).
Step 0: Of course if Pic R is trivial then D(Pic R) = 1, ρ(R) = 1 and the result
holds in this case. Henceforth we assume Pic R is nontrivial and thus D(Pic R) ≥ 2,
and our task is to show that ρ(R) ≤ D(Pic2
R)
, with equality if R is replete.
• ×
Step 1: Let x ∈ R \ R . Consider two irreducible factorizations
x = α1 · · · αm = β1 · · · βn
of x with m ≥ n. Let k be the number of principal prime ideals in the prime
ideal factorization of (x). Then k = n ⇐⇒ k = m =⇒ ρ(x) = 1. Henceforth
we assume k < min(m, n) (since Pic R is nontrivial, there is at least one such
x). We may further assume that α1 , . . . , αk (resp. β1 , . . . , βk ) are prime elements
and αk+1 , . . . , αm (resp. βk+1 , . . . , βn ) are not; dividing through by these prime
elements and correcting by a unit if necessary, we may write
x0 = αk+1 · · · αm = βk+1 · · · βn .
Since for k + 1 ≤ i ≤ m, αi is irreducible but not prime, P (αi ) ≥ 2 and thus
2(m − k) ≤ P (αk+1 · · · αm ) = P (x0 ).
On the other hand, by Proposition 23.13 we have
P (x0 ) = P (βk+1 · · · βn ) ≤ (n − k)D(Pic R).
Combining these inequalities gives
m m−k D(Pic R)
≤ ≤ .
n n−k 2
It follows that ρ(x) ≤ D(Pic 2
R)
and thus ρ(R) ≤ D(Pic
2
R)
, establishing part a).
Step 2: Suppose R is replete.
Step 2a: Suppose first that Pic R is finite and put D = D(Pic R). By repleteness,
choose prime ideals p1 , . . . , pD whose classes form an irreducible sequence in Pic R.
For 1 ≤ i ≤ D, let qi be a prime ideal with [qi ] = [pi ]−1 . For 1 ≤ i ≤ D, let

1Here we are considering G as an additive group.


372 23. STRUCTURE OF OVERRINGS

ci be such that (ci ) = pi qi ; using Lemma 23.11 there are d1 , d2 ∈ R such that
(d1 ) = p1 · · · pD and (d2 ) = q1 · · · qD and
c1 · · · cD = d1 d2 .
By Proposition 23.12, c1 , . . . , cD , d1 , d2 are all irreducible, and thus ρ(R) ≥ D
2.
Step 2b: If Pic R is infinite, then D(Pic R) = ∞ and from this, repleteness and
Lemma 23.11, for all D ∈ Z+ there are prime ideals p1 , . . . , pD whose classes form
an irreducible sequence in Pic R and such that p1 · · · pD is principal. The argument
of Step 2a now shows ρ(R) ≥ D +
2 . Since this holds for all D ∈ Z , ρ(R) = ∞. 
Remark: When Pic R is finite, Theorem 23.15 is due to W. Narkiewicz [Na95].

Remark: The condition that R be replete is essential in Theorem 23.15. For in-
stance, A. Zaks has shown that for every finitely generated commutative group G,
there is a half factorial Dedekind domain R with Pic R ∼= G [Za76]. Whether any
commutative group can occur, up to isomorphism, as the Picard group of a half
factorial Dedekind domain is an open problem.
Corollary 23.16. a) A replete Dedekind domain R is a HFD iff # Pic R ≤ 2.
b) (Carlitz [Ca60]) Let K be a number field. Then its ring of integers ZK is a HFD
iff the class number of K – i.e., # Pic ZK – is either 1 or 2.
c) (Valenza [Va90]) Let K be a number field. Then
 
D(Pic ZK )
ρ(ZK ) = max ,1 .
2
d) A replete Dedekind domain has infinite elasticity iff it has infinite Picard group.
Exercise 23.8. Prove Corollary 23.16.
Later we will show that every commutative group arises, up to isomorphism, as
the Picard group of a Dedekind domain. Combining this with Theorems 23.10 and
23.15 we see that the possible elasticities for replete Dedekind domains are precisely
n
2 for any integer n ≥ 2 and ∞.

We end this section by giving a little more information on the Davenport con-
stant: let G be a finite commutative group, so that there is a uniqueLsequence of
positive integers n1 , . . . , nr with nr | nr−1 | . . . | n1 > 1 such that G ∼
r
= i=1 Z/ni Z.
Pk
We put d(G) = 1 + i=1 (ni − 1), so that (49) reads more succinctly as
(50) D(G) ≥ d(G).
J.E. Olson conjectured that equality holds in (49) for all finite commutative groups
G [Ol69a]. He proved that the conjecture holds for p-groups [Ol69a] and also when
r ≤ 2 [Ol69b]. However, it was shown by P. van Emde Boas and D. Kruyswijk
that (for instance) D(G) > d(G) for G = Z/6Z × Z/3Z × Z/3Z × Z/3Z [EBK69].
Whether D(G) = d(G) for all groups with r = 3 is still an open problem. In
particular the exact value of D(G) is unknown for most finite commutative groups.

4. Overrings of Prüfer Domains


Theorem 23.17. Let R be a domain with fraction field K, and let T be an
overring of R. the following are equivalent:
(i) For every prime ideal p of R, either pT = T or T ⊂ Rp .
4. OVERRINGS OF PRÜFER DOMAINS 373

(ii) For every x, y ∈ K × with x


y ∈ T , ((y) : (x))T = T .
(iii) T is a flat R-algebra.
Proof. COMPLETE ME! 

Proposition 23.18. Let R be a domain with fraction field K and consider


rings R ⊂ T ⊂ T 0 ⊂ K.
a) If T 0 is flat over R, then T 0 is flat over T .
b) If T 0 is flat over T and T is flat over R, then T 0 is flat over R.
Proof. a) Suppose T 0 is flat over R. Let a, b ∈ T be such that ab ∈ T 0 . Write
a = sc , b = ds with c, d, s ∈ R. Then dc ∈ T 0 , so by Theorem 23.17, ((d) : (c))T 0 = T 0 .
Hence 1 = t1 u1 + . . . + tk uk for some ti ∈ T 0 and ui ∈ R with ui c ∈ (d) for all i.
Then there is zi ∈ R such that ui c = dzi , so ui a = zi ds = zi b ∈ T b for all i. So
(T b : T a)T 0 = T 0 . Applying Theorem 23.17 again, we get that T 0 is flat over T .
b) This holds for any R1 ⊂ R2 ⊂ R3 , since M ⊗R1 R3 ∼ = (M ⊗R1 R2 ) ⊗R2 R3 . 

Proposition 23.19. For an overring T of a domain R, the following are equiv-


alent:
(i) T is flat over R.
T p ∈ MaxSpec T , Tp = Rp∩R .
(ii) For all
(iii) T = p∈MaxSpec T Rp .
Proof. COMPLETE ME! 

Proposition 23.20. Let T be an overring of a domain R which is both integral


and flat over R. Then R = T .
Proof. Let x, y ∈ R be such that xy ∈ T . Then by Theorem 23.17, ((y) :
(x))T = T . Let p ∈ MaxSpec R. By Theorem 14.13, there exists a prime (in fact
maximal by Corollary 14.16, but this is not needed here) ideal P of T lying over
p. Since pT ⊂ P, we have pT ( T . Therefore ((y) : (x)) is not contained in any
maximal ideal of R, so ((y) : (x)) = R. It follows that x ∈ (y), i.e., x = ay for some
a ∈ R, so that xy ∈ R. Thus R = T . 

Theorem 23.21. For a domain R, the following are equivalent:


(i) R is a Prüfer domain.
(ii) Every overring of R is a flat R-module.
Proof. (i) =⇒ (ii): An overring is a torsionfree R-module, so this follows
immediately from Theorem 21.10.
(ii) =⇒ (i): Let p be a maximal ideal of R. By Proposition 23.18, every overring
of Rp is flat. Let a, b ∈ Rp , and suppose that aRp 6⊂ bRp . Then (bRp : aRp ) 6= Rp ;
since Rp is local, this implies (bRp : aRp ) ⊂ pRp . Now consider the ring Rp [ ab ].
This is an overring of Rp , hence flat. Since ab ∈ Rp [ ab ] we have by Theorem 23.17
a a
(bRp : aRp )Rp [ ] = Rp [ ].
b b
Thus there exist elements x1 , . . . , xn ∈ (bRp : aRp and b1 , . . . , bn ∈ Rp [ ab ] such that
x1 b1 + . . . + xn bn = 1. . . . COMPLETE ME! 

Corollary 23.22. Every overring of a Prüfer domain is a Prüfer domain.


374 23. STRUCTURE OF OVERRINGS

Proof. Let R be a Prüfer domain and T be an overring of R. Then every


overring T 0 of T is in particular an overring of R, so T 0 is flat over R. By Proposition
23.18a), T 0 is also flat over T . Therefore every overring of T is flat over T , so by
Theorem 23.21, T is a Prüfer domain. 
Corollary 23.23. For a Noetherian domain R, the following are equivalent:
(i) Every overring of R is a localization.
(ii) R is a Dedekind domain and Pic R is a torsion group.
Exercise 23.9. Prove Corollary 23.23.
Finally we give a result which generalizes the (as yet unproven) Theorem 23.3.
Namely, for R a domain, let W be a subset of MaxSpec R and put
\
RW = Rp .
p∈W

Theorem 23.24. Let R be a Prüfer domain, T an overring of R, and put


W = {p ∈ MaxSpec(R) | pT ( T }.
Then T = RW .
Proof. COMPLETE ME! 

5. Kaplansky’s Theorem (III)


Theorem 23.25. (Kaplansky [K]) Let R be a Dedekind domain with fraction
field K, and let K be an algebraic closure of K. Suppose that for every finite
extension L/K, the Picard group of the integral closure RL of R in L is a torsion
commutative group. Then the integral closure S of R in K is a Bézout domain.
Proof. Let I = ha1 , . . . , an i be a finitely generated ideal of S. Then L =
K[a1 , . . . , an ] is a finite extension of K. Let RL be the integral closure of R in
L, and let IL = ha1 , . . . , an iRL . By hypothesis, there exists k ∈ Z+ and b ∈ RL
such that ILk = bRL . Let c be a kth root of b in S and let M = L[c]. Thus in the
Dedekind domain RM we have (IL RM )k = (ck ), and from unique factorization of
ideals we deduce IL RM = cRM . Thus I = IL RM S = cRM S = cS is principal. 
Recall the basic fact of algebraic number theory that for any number field K, the
Picard group of ZK is finite. This shows that the ring R = Z satisfies the hypotheses
of Theorem 23.25. We deduce that the ring of all algebraic integers Z is a Bézout
domain: Theorem 5.1.
Exercise 23.10. Adapt the proof of Theorem 23.25 to show that the Picard
group of the ring of integers of the maximal solvable extension Qsolv of Q is trivial.
Exercise 23.11. State a function field analogue of Theorem 5.1 and deduce it
as a special case of Theorem 23.25.
We quote without proof two more results on Picard groups of integer rings of infinite
algebraic extensions of Q.
Theorem 23.26. (Brumer [Br81]) Let Qcyc = n∈Z+ Q(ζn ) be the field ob-
S
tained by adjoining to Q all roots of unity, and let Zcyc be its ring of integers, i.e.,
the integral closure of Z in Qcyc . Then

Pic Zcyc ∼
M
= Q/Z.
i=1
6. EVERY COMMUTATIVE GROUP IS A CLASS GROUP 375

Theorem 23.27. (Kurihara [Ku99]) Let Qcyc + = n∈Z+ Q(ζn + ζn−1 ) be the
S
maximal real subfield of Qcyc , and let Zcyc + be its ring of integers, i.e., the integral
closure of Z in Qcyc + . Then
Pic Zcyc + = 0.

6. Every commutative group is a class group


To any ring R we attached a commutative group, the Picard group Pic R. In fact
the construction is functorial: a homomorphism ϕ : R → R0 of domains induces a
homomorphism ϕ∗ : Pic R → Pic S of Picard groups. Explicitly, if M is a rank one
projective R-module, then M ⊗R R0 is a rank one projective R0 -module. In general
when one is given a functor it is natural to ask about its image. Here we are asking
the following
Question 6. Which commutative groups occur (up to isomorphism) as the
Picard group of a commutative ring?
We have also defined the divisor class group Cl R of a domain, so we may also ask:
Question 7. Which commutative groups occur (up to isomorphism) as the
divisor class group of a domain?
It would be interesting to know at what point algebraists began serious consider-
ation of the above questions. I have not discussed the history of Pic R and Cl R
in part because I am not sufficiently knowledgeable to do so, but their study was
surely informed by two classical cases: the ideal class group of (the ring of integers
of) a number field, and the Picard group of line bundles on an (affine or projective)
complex algebraic variety. The groups that arise in these classical cases are very
restricted: the class group of a number field is a finite commutative group, and the
Picard group of a complex algebraic variety is an extension of a complex torus by
a finitely generated commutative group. As far as I know the literature contains
nothing beyond this until the dramatic full solution.
Theorem 23.28. (Claborn [Cl66]) For every commutative group G, there is a
Dedekind domain R with Pic R = Cl R ∼
= G.
Even after developing several hundred pages of commutative algebra, Claborn’s
proof still requires some technical tools that we lack. Especially, Claborn first con-
structs a Krull domain R with Cl R ∼ = G and then by an approximation process
constructs a Dedekind domain with the same class group. But we have not yet
discussed Krull domains in these notes.
A more elementary – but still quite ingenious and intricate – proof was given
later by C.R. Leedham-Green [Le72]. Leedham-Green constructs the requisite R
as the integral closure of a PID in a separable quadratic field extension.
Several years after that M. Rosen took a more naturally geometric approach,
inspired by the Picard groups of varieties which appear in algebraic geometry. Es-
pecially, his approach uses some elliptic curve theory.

Let k be a field of characteristic zero. Fix elements A, B ∈ K such that 4A3 +


27B 2 6= 0, and let k[E] = k[x, y]/(y 2 − x3 − Ax − B). The ring k[E] is precisely the
affine coordinate ring of the elliptic curve
E : y 2 = x3 + Ax + B.
376 23. STRUCTURE OF OVERRINGS

Proposition 23.29. The ring k[E] is a Dedekind domain.


Exercise 23.12. Prove Proposition 23.29. (Suggestions: it should be clear that
k[E] is a one-dimensional Noetherian domain, so the matter of it is to show that it
is integrally closed. For this, show that the condition that 4A3 + 27B 2 6= 0 means
that x3 − Ax − B ∈ k[x] is separable – i.e., has distinct roots in an algebraic closure
– and thus is squarefree. Apply Theorem 15.17.)
We denote the fraction field of k[E] by k(E), also called the function field of E/k .

The overrings of R are all of the form RW = p∈ΣR \W Rp for W ⊂ MaxSpec R,


T

and by the Krull-Akizuki Theorem each RW is a Dedekind domain. By definition,


a Dedekind domain which arises in this way – i.e., as an overring of the standard
affine ring of an elliptic curve over a field of characteristic zero – is called an elliptic
Dedekind domain. We refer to k as the ground field of R.
Exercise 23.13. Let k be a countable field, and let R be an elliptic Dedekind
domain with ground field k. Show: Pic R is countable.
b) More generally, show: if R is an elliptic Dedekind domain with ground field k,
then # Pic R ≤ max ℵ0 , #k.
Conversely:
Theorem 23.30. (Rosen [Ro76]) For every countable commutative group G,
there exists an elliptic Dedekind domain R with ground field an algebraic extension
of Q and such that Pic R ∼ = G.
In 2008 I built on this work of Rosen to prove the following result [Cl09].
Theorem 23.31. For any commutative group G, there is an elliptic Dedekind
domain R such that:
(i) R is the integral closure of a PID in a separable quadratic field extension,
and
(ii) Pic R ∼
= G.
Thus Theorem 23.31 implies the results of Claborn and Leedham-Green. On the
other hand, Exercise 23.12 shows that the absolute algebraicity (or even the count-
ability!) of the ground field k achieved in Rosen’s construction cannot be maintained
for uncountable Picard groups. Indeed our argument goes to the other extreme:
we construct the ground field k as a transfinitely iterated function field.

Our argument will require some tenets of elliptic curve theory, especially the no-
tion of the rational endomorphism ring Endk E of an elliptic curve E/k . A
k-rational endomorphism of an elliptic curve is a morphism ϕ : E → E defined over
k which carries the neutral point O of E to itself.
Proposition 23.32. Let k be a field and E/k an elliptic curve.
a) The additive group of Endk (E) is isomorphic to Za(E) for a(E) ∈ {1, 2, 4}.
b) There is a short exact sequence
0 → E(k) → E(k(E)) → Endk (E) → 0.
Since Endk (E) is free abelian, we have E(k(E)) ∼
= E(k) Za(E) .
L
c) There is a canonical isomorphism E(k) ∼= Pic k[E].
6. EVERY COMMUTATIVE GROUP IS A CLASS GROUP 377

Proof. a) See [Si86, Cor. III.9.4].


b) E(k(E)) is the group of rational maps from the nonsingular curve E to the
complete variety E under pointwise addition. Every rational map from a nonsin-
gular curve to a complete variety is everywhere defined, so E(k(E)) is the group
of morphisms E → E under pointwise addition. The constant morphisms form a
subgroup isomorphic to E(k), and every map E → E differs by a unique constant
from a map of elliptic curves (E, O) → (E, O), i.e., an endomorphism of E.
c) By Riemann-Roch, Ψ1 : E(k) → Pic0 E by P ∈ E(k) 7→ [[P ] − [O]] is an
isomorphism [Si86, Prop. III.3.4]. Moreover, Ψ2 : Pic0 (E) → Pic k[E] given by
P P ∼
P nP [P ] 7→ P 6=O nP [P ] is an isomorphism. Thus Ψ2 ◦Ψ1 : E(k) → Pic k[E]. 

Now fix a field k, and let (E0 )/k be any elliptic curve.2 Define K0 = k, and
Kn+1 = Kn (E/Kn ). Then Proposition 23.32 gives
n
E(Kn ) ∼
M
= E(k) ⊕ Za(E) .
i=1

Lemma 23.33. Let K be a field, (Ki )i∈I a directed system of field extensions
of K, and E/K an elliptic curve. There is a canonical isomorphism
lim E(Ki ) = E(lim Ki ).
I I

Exercise 23.14. Prove Lemma 23.33.


Now let o be an ordinal number. We define the field Ko by transfinite induction:
K0 = k, for an ordinal o0 < o, Ko0 +1 = Ko0 (E/Ko0 ), and for a limit ordinal o,
Ko = limo0 <o Ko0 . By the Continuity Lemma, we have E(Ko ) = limo0 ∈o E(Ko0 ).
Lemma 23.34. Let a ∈ Z+ . For a commutative group A, the following are
equivalent:
(i) A is free abelian of rank a · κ for some cardinal κ.
(ii) A has a well-ordered ascending series with all factors As+1 /As ∼
= Za .
Exercise 23.15. Prove Lemma 23.34.
(Suggestion: use the Transfinite Dévissage Lemma.)
Corollary 23.35. We have E(Ko )/E(k) ∼ = o0 ∈o Za(E) .
L

Exercise 23.16. Prove Corollary 23.35.


One can put together the results derived so far together with Exercise 22.2 to get a
proof of Theorem 23.28. However, to prove Theorem 23.31 we need to circumvent
the appeal to Theorem 23.10. This is handled as follows.
Theorem 23.36. Let E/k be an elliptic curve with equation y 2 = P (x) =
x3 + Ax + B. a) The affine ring k[E] is weakly replete.
b) If k is algebraically closed, k[E] is not replete.
c) Suppose k does not have characteristic 2 and k[E] is not replete. Then for all
x ∈ k, there exists y ∈ k with y 2 = P (x).
2In the paper [Cl09] I took the specific choice (E ) 2 3
0 /Q : y + y = x − 49x − 86, mostly
for sentimental reasons. This curve has the property that E0 (Q) = 0 [Ko89, Theorem H] and
12 3
nonintegral j-invariant 2 373 , so EndQ E = Z. But in fact the construction can be made to work
starting with any elliptic curve, and we have decided to phrase it this way here.
378 23. STRUCTURE OF OVERRINGS

Proof. Each point P 6= O on E(k) a prime ideal in the standard affine ring
k[E]; according to the isomorphism of Proposition 23.32c), every nontrivial element
of Pic(k[E]) arises in this way. This proves part a). Part b) is similar: if k is
algebraically closed, then by Hilbert’s Nullstellensatz every prime ideal of k[E]
corresponds to a k-valued point P 6= O on E(k), which under Proposition 23.32c)
corresponds to a nontrivial element of the class group. Therefore the trivial class is
not represented by any prime ideal. p Under the hypotheses of part c), there exists
an x ∈ k such that
p the points (x,
p ± P (x)) form a Galois conjugate pair. Therefore
the divisor (x, P (x)) + (x, − P (x)) represents a closed point on the curve C o ,
in other
p words a nonzero
p prime ideal of k[E]. But the corresponding point on E(k)
is (x, P (x)) + (x, − P (x)) = O. 
Finally we prove Theorem 23.31(i). Let G be a commutative group, and write it
as F/H where F is a free commutative group of infinite rank. As above let k be
any field of characteristic zero and E/k any elliptic curve. By Corollary 23.35, for
all sufficiently large ordinals o, there is a surjection E(K0 ) → F and thus also a
surjection E(Ko ) → G. By Proposition 23.32c), there is a subgroup H of Ko [E]
such that (Pic Ko [E])/H ∼ = G. By Proposition 18.1a) and Proposition 23.8, there
is an overring T of Ko [E] such that Pic T ∼ = G, establishing Theorem 23.31(i).

As for the second part: let σ be the automorphism of the function field k(E)
induced by (x, y) 7→ (x, −y), and notice that σ corresponds to inversion P 7→ −P
on E(k) = Pic(k[E]). Let S = Rσ be the subring of R consisting of all functions
which are fixed by σ. Then k[E]σ = k[x] is a PID, and S is an overring of k[x],
hence also a PID. More precisely, S is the overring of all functions on the projective
line which are regular away from the point at infinity and the x-coordinates of all
the elements in H (since H is a subgroup, it is stable under inversion). Finally,
to see that R is the integral closure of S in the separable quadratic field extension
k(E)/k(x), it suffices to establish the following simple result.
Lemma 23.37. Let L/K be a finite Galois extension of fields, and S a Dedekind
domain with fraction field L. Suppose that for all σ ∈ Gal(L/K), σ(S) = S. Then
S is the integral closure of R := S ∩ K in L.
Proof. Since S is integrally closed, it certainly
Q contains the integral closure
of R in L. Conversely, for any x ∈ S, P (t) = σ∈Gal(L/K) (t − σ(x)) is a monic
polynomial with coefficients in (S ∩ K)[t] satisfied by x. 
This completes the proof of Theorem 23.31.
Bibliography

[AB59] M. Auslander and D.A. Buchsbaum, Unique factorization in regular local rings. Proc.
Nat. Acad. Sci. U.S.A. 45 (1959), 733–734.
[Ad62] J.F. Adams, Vector fields on spheres. Ann. of Math. (2) 75 (1962), 603–632.
[AHRT] J. Adámek, H. Herrlich, J. Rosický and W. Tholen, Injective hulls are not natural.
Algebra Universalis 48 (2002), 379–388.
[AKP98] D.D. Anderson, B.G. Kang and M.H. Park, Anti-Archimedean rings and power series
rings. Comm. Algebra 26 (1998), 3223—3238.
[Al63] N.L. Alling, The valuation theory of meromorphic function fields over open Riemann
surfaces. Acta Math. 110 (1963), 79–96.
[Al99] N. Alon, Combinatorial Nullstellensatz. Recent trends in combinatorics (Mátraháza,
1995). Combin. Probab. Comput. 8 (1999), 7–29.
[An00] D.D. Anderson, GCD domains, Gauss’ lemma, and contents of polynomials. Non-
Noetherian commutative ring theory, 1–31, Math. Appl., 520, Kluwer Acad. Publ.,
Dordrecht, 2000.
[AR97] D.D. Anderson and M. Roitman, A characterization of cancellation ideals. Proc.
Amer. Math. Soc. 125 (1997), 2853–2854.
[AZ94] D.D. Anderson and M. Zafrullah, On a theorem of Kaplansky. Boll. Un. Mat. Ital. A
(7) 8 (1994), 397–402.
[AP82] J.K. Arason and A. Pfister, Quadratische Formen über affinen Algebren und ein al-
gebraischer Beweis des Satzes von Borsuk-Ulam. (German) J. Reine Angew. Math.
331 (1982), 181–184.
[Ar27] E. Artin, Zur Theorie der hyperkomplexen Zahlen. Abh. Hamburg, 5 (1927), 251-260.
[AT51] E. Artin and J.T. Tate, A note on finite ring extensions. J. Math. Soc. Japan 3 (1951),
74–77.
[At89] M.F. Atiyah, K-theory. Notes by D. W. Anderson. Second edition. Advanced Book
Classics. Addison-Wesley Publishing Company, Redwood City, CA, 1989.
[AM] M.F. Atiyah and I.G. Macdonald, Introduction to commutative algebra. Addison-
Wesley Publishing Co., Reading Mass.-London-Don Mills, Ont. 1969.
[Ba40] R. Baer, Abelian groups that are direct summands of every containing commutative
group. Bull. Amer. Math. Soc. 46 (1940), 800–806.
[Ba59] H. Bass, Global dimension of rings, Ph.D. Thesis, University of Chicago, 1959.
[Ba63] H. Bass, Big projective modules are free. Illinois J. Math. 7 1963 24–31.
[Ba62] G. Baumslag, On abelian hopfian groups. I. Math. Z. 78 1962 53–54.
[Ba63] G. Baumslag, Hopficity and commutative groups. 1963 Topics in Abelian Groups
(Proc. Sympos., New Mexico State Univ., 1962) pp. 331–335 Scott, Foresman and
Co., Chicago, Ill.
[BCR] J.Bochnak, M. Coste and M.-F. Roy, Real algebraic geometry. Ergebnisse der Mathe-
matik und ihrer Grenzgebiete (3) 36. Springer-Verlag, Berlin, 1998.
[Be] D.J. Benson, Polynomial invariants of finite groups London Mathematical Society
Lecture Note Series, 190. Cambridge University Press, Cambridge, 1993.
[Bo50] S. Borofsky, Factorization of polynomials. Amer. Math. Monthly 57 (1950), 317–320.
[BM58] R. Bott and J. Milnor, On the parallelizability of the spheres. Bull. Amer. Math. Soc.
64 (1958), 87–89.
[B] N. Bourbaki, Commutative algebra. Chapters 1–7. Translated from the French.
Reprint of the 1989 English translation. Elements of Mathematics (Berlin). Springer-
Verlag, Berlin, 1998.

379
380 BIBLIOGRAPHY

[Bo78] A. Bouvier, Le groupe des classes de l’algebre affine d’une forme quadratique. Publ.
Dép. Math. (Lyon) 15 (1978), no. 3, 53–62.
[BMRH] J.W. Brewer, P.R. Montgomery, E.A. Rutter and W.J. Heinzer, Krull dimension
of polynomial rings. Conference on Commutative Algebra (Univ. Kansas, Lawrence,
Kan., 1972), pp. 26-45. Lecture Notes in Math., Vol. 311, Springer, Berlin, 1973.
[Br02] G. Brookfield, The length of Noetherian modules. Comm. Algebra 30 (2002), 3177–
3204.
[Br81] A. Brumer, The class group of all cyclotomic integers. J. Pure Appl. Algebra 20
(1981), no. 2, 107–111.
[Bu61] D.A. Buchsbaum, Some remarks on factorization in power series rings. J. Math.
Mech. 10 1961 749–753.
[Bu64] H.S. Butts, Unique factorization of ideals into nonfactorable ideals. Proc. Amer. Math.
Soc. 15 (1964), 21.
[BW66] H.S. Butts and L.I. Wade, Two criteria for Dedekind domains. Amer. Math. Monthly
73 (1966), 14–21.
[Ca60] L. Carlitz, A Characterization of Algebraic Number Fields with Class Number Two.
Proc. AMS 11 (1960), 391–392.
[Cd-SF] K. Conrad, Stably Free Modules. Notes available at
http://www.math.uconn.edu/∼kconrad/blurbs/linmultialg/stablyfree.pdf
[CDVM13] L.F. Cáceres Duque and J.A. Vélez-Marulanda, On the Infinitude of Prime Elmeents.
Rev. Colombiana Mat. 47 (2013), 167–179.
[CE] H. Cartan and S. Eilenberg, Homological algebra. Princeton University Press, Prince-
ton, N. J., 1956.
[CE59] E.D. Cashwell and C.J. Everett, The ring of number-theoretic functions. Pacific J.
Math. 9 (1959) 975–985.
[CK46] I.S. Cohen and I. Kaplansky, Rings with a finite number of primes. I. Trans. Amer.
Math. Soc. 60 (1946), 468–477.
[CK51] I.S. Cohen and I. Kaplansky, Rings for which every module is a direct sum of cyclic
modules. Math. Z. 54 (1951), 97–101.
[Cl66] L.E. Claborn, Every commutative group is a class group. Pacific J. Math. 18 (1966),
219–222.
[Cl-GT] P.L. Clark, General Topology. http://math.uga.edu/~pete/pointset2018.pdf.
[Cl09] P.L. Clark, Elliptic Dedekind domains revisited. Enseignement Math. 55 (2009), 213–
225.
[Cl15] P.L. Clark, A note on Euclidean order types. Order 32 (2015), 157–178.
[Cl17a] P.L. Clark, The Euclidean criterion for irreducibles. Amer. Math. Monthly 124 (2017),
198–216.
[Cl17b] P.L. Clark, The cardinal Krull dimension of a ring of holomorphic functions. Expo.
Math. 35 (2017), 350–356.
[Cl18] P.L. Clark, A note on rings of finite rank. Comm. Algebra 46 (2018), 4223–4232.
[Cn68] P.M. Cohn, Bézout rings and their subrings. Proc. Cambridge Philos. Soc. 64 (1968),
251–264.
[Cn73] P.M. Cohn, Unique factorization domains. Amer. Math. Monthly 80 (1973), 1–18.
[Co46] I.S. Cohen, On the structure and ideal theory of complete local rings. Trans. Amer.
Math. Soc. 59, (1946), 54–106.
[Co50] I.S. Cohen, Commutative rings with restricted minimum conditions. Duke Math. J.
17, (1950), 27–42.
[CS46] I.S. Cohen and A. Seidenberg, Prime ideals and integral dependence. Bull. Amer.
Math. Soc. 52 (1946), 252–261.
[Co64] A.L.S. Corner, On a conjecture of Pierce concerning direct decompositions of Abelian
groups. 1964 Proc. Colloq. Abelian Groups (Tihany, 1963) pp. 43–48 Akadémiai Kiadó,
Budapest.
[Co11] D.A. Cox, Why Eisenstein proved the Eisenstein criterion and why Schönemann dis-
covered it first. Amer. Math. Monthly 118 (2011), 3–21.
[Da09] C.S. Dalawat, Wilson’s theorem. J. Théor. Nombres Bordeaux 21 (2009), 517–521.
[De71] R. Dedekind. . .
[DM71] F. DeMeyer and E. Ingraham, Separable algebras over commutative rings. Lecture
Notes in Mathematics, Vol. 181 Springer-Verlag, Berlin-New York 1971.
BIBLIOGRAPHY 381

[Ea68] P.M. Eakin, Jr. The converse to a well known theorem on Noetherian rings. Math.
Ann. 177 (1968), 278–282.
[EBK69] P. van Emde Boas and D. Kruyswijk, A combinatorial problem on finite commutative
groups, III, Report ZW-1969-008, Math. Centre, Amsterdam, 1969.
[Ec00] O. Echi, A topological characterization of the Goldman prime spectrum of a commu-
tative ring. Comm. Algebra 28 (2000), 2329–2337.
[ES53] B. Eckmann and A. Schopf, Über injektive Moduln. Arch. Math. (Basel) 4 (1953),
75–78.
[Ei] D. Eisenbud, Commutative algebra. With a view toward algebraic geometry. Graduate
Texts in Mathematics, 150. Springer-Verlag, New York, 1995.
[Ei50] F. Eisenstein, Über die Irredicibilität une einige andere Eigenschaften der Gleichung
von welche der Theilung der ganzen Lemniscate abhängt. J. Reine Angew. Math. 39
(1950), 160–179.
[Fo64] O. Forster, Über die Anzahl der Erzeugenden eines Ideals in einem Noetherschen
Ring. Math. Z. 84 (1) (1964) 80-–87.
[FR70] R.L. Finney and J.J. Rotman, Paracompactness of locally compact Hausdorff spaces.
Michigan Math. J. 17 (1970), 359–361.
[FT] Field Theory, notes by P.L. Clark, available at
http://www.math.uga.edu/∼pete/FieldTheory.pdf
[FW67] C. Faith and E.A. Walker, Direct-sum representations of injective modules. J. Algebra
5 (1967), 203–221.
[Fl71] C.R. Fletcher, Euclidean rings. J. London Math. Soc. 4 (1971), 79–82.
[Fo73] E. Formanek, Faithful Noetherian modules. Proc. Amer. Math. Soc. 41 (1973), 381–
383.
[Fs73] R.M. Fossum, The divisor class group of a Krull domain. Ergebnisse der Mathematik
und ihrer Grenzgebiete, Band 74. Springer-Verlag, New York-Heidelberg, 1973.
[Fu59] L. Fuchs, The existence of indecomposable commutative groups of arbitrary power.
Acta Math. Acad. Sci. Hungar. 10 (1959) 453–457.
[Fu74] L. Fuchs, Indecomposable commutative groups of measurable cardinalities. Symposia
Mathematica, Vol. XIII (Convegno di Gruppi Abeliani, INDAM, Rome, 1972), pp.
233-244. Academic Press, London, 1974.
[Ga71] T.E. Gantner, A Regular Space on which every Continuous Real-Valued Function is
Constant. Amer. Math. Monthly 78 (1971), 52–53.
[Ge91] R. Germundsson, Basic Results on Ideals and Varieties in Finite Fields. Technical
report, Department of Electrical Engineering, Lı̈nkoping University, 1991.
[Gi72] R. Gilmer, On commutative rings of finite rank. Duke Math. J. 39 (1972), 381—383.
[GJ76] L. Gillman and M. Jerison, Rings of continuous functions. Reprint of the 1960 edition.
Graduate Texts in Mathematics, No. 43. Springer-Verlag, New York-Heidelberg, 1976.
[Gol51] O. Goldman, Hilbert rings and the Hilbert Nullstellensatz. Math. Z. 54 (1951). 136–
140.
[Gol64] O. Goldman, On a special class of Dedekind domains. Topology 3 (1964) suppl. 1,
113–118.
[GoN] P.L. Clark, Geometry of numbers with applications to number theory. http://math.
uga.edu/~pete/geometryofnumbers.pdf
[Gov65] V.E. Govorov, On flat modules. (Russian) Sibirsk. Mat. Ž. 6 (1965), 300–304.
[GoRo] R. Gordon and J.C. Robson, Krull dimension. Memoirs of the American Mathematical
Society, No. 133. American Mathematical Society, Providence, R.I., 1973.
[GuRo] R.C. Gunning and H. Rossi, Analytic functions of several complex variables. Prentice-
Hall, Inc., Englewood Cliffs, N.J. 1965.
[Gr74] A. Grams, Atomic rings and the ascending chain condition for principal ideals. Proc.
Cambridge Philos. Soc. 75 (1974), 321–329.
[Gu73] T.H. Gulliksen, A theory of length for Noetherian modules. J. Pure Appl. Algebra 3
(1973), 159–170.
[Ha] R. Hartshorne, Algebraic geometry. Graduate Texts in Mathematics, No. 52. Springer-
Verlag, New York-Heidelberg, 1977.
[Ha94] B. Haible, Gauss’ Lemma without Primes, preprint, 1994.
[Ha28] H. Hasse, Über eindeutige Zerlegung in Primelemente oder in Primhauptideale in
Integrit-tsbereichen. J. reine Angew. Math. 159 (1928), 3–12.
382 BIBLIOGRAPHY

[He40] O. Helmer, Divisibility properties of integral functions. Duke Math. J. 6 (1940), 345–
356.
[He53] M. Henriksen, On the prime ideals of the ring of entire functions. Pacific J. Math. 3
(1953), 711–720.
[He70] W. Heinzer, Quotient overrings of an integral domain. Mathematika 17 (1970), 139–
148.
[He74] R.C. Heitmann, PID’s with specified residue fields. Duke Math. J. 41 (1974), 565–582.
[He06] G. Hessenberg, Grundbegriffe der Mengenlehre. Göttingen, 1906.
[Hi90] D. Hilbert, Ueber die Theorie der algebraischen Formen. Mathl Annalen 36 (1890),
473–534.
[Ho69] M. Hochster, Prime ideal structure in commutative rings. Trans. Amer. Math. Soc.
142 (1969), 43–60.
[Hi75] J.-J. Hiblot, Des anneaux euclidiens dont le plus petit algorithme n’est pas à valeurs
finies. C. R. Acad. Sci. Paris Sér. A-B 281 (1975), no. 12, Ai, A411–A414.
[Hi77] J.-J. Hiblot, Correction à une note sur les anneaux euclidiens: “Des anneaux eu-
clidiens dont le plus petit algorithme n’est pas à valeurs finies” (C. R. Acad. Sci.
Paris Sér. A-B 281 (1975), no. 12, A411–A414). C. R. Acad. Sci. Paris S er. A-B
284 (1977), no. 15, A847–A849.
[Hö01] O. Hölder, Die Axiome der Quantitat und die Lehre vom Mass. Ber. Verh. Sachs. Ges.
Wiss. Leipzig, Math.-Phys. Cl. 53 (1901), 1-64.
[Ho79] W. Hodges, Krull implies Zorn. J. London Math. Soc. (2) 19 (1979), 285–287.
[Hu68] T.W. Hungerford, On the structure of principal ideal rings. Pacific J. Math. 25 (1968),
543–547.
[Hs66] D. Husemöller, Fibre bundles. McGraw-Hill Book Co., New York-London-Sydney 1966.
[J1] N. Jacobson, Basic algebra. I. Second edition. W. H. Freeman and Company, New
York, 1985.
[J2] N. Jacobson, Basic algebra. II. Second edition. W. H. Freeman and Company, New
York, 1989.
[Jo00] P. Jothilingam, Cohen’s theorem and Eakin-Nagata theorem revisited. Comm. Algebra
28 (2000), 4861–4866.
[JR80] M. Jarden and P. Roquette, The Nullstellensatz over p-adically closed fields. J. Math.
Soc. Japan 32 (1980), 425–460.
[Ka49] I. Kaplansky, Elementary divisors and modules. Trans. Amer. Math. Soc. 66 (1949),
464–491.
[Ka52] I. Kaplansky, Modules over Dedekind rings and valuation rings. Trans. Amer. Math.
Soc. 72 (1952), 327–340.
[Ka58] I. Kaplansky, Projective modules. Ann. of Math. 68 (1958), 372–377.
[K] I. Kaplansky, Commutative rings. Allyn and Bacon, Inc., Boston, Mass. 1970.
[Ka] M. Karoubi, K-theory. An introduction. Reprint of the 1978 edition. With a new
postface by the author and a list of errata. Classics in Mathematics. Springer-Verlag,
Berlin, 2008.
[KeOm10] K.A. Kearnes and G. Oman, Cardinalities of residue fields of Noetherian integral
domains. Comm. Algebra 38 (2010), 3580–3588.
[Kh03] D. Khurana, On GCD and LCM in domains - a conjecture of Gauss. Resonance 8
(2003), 72–79.
[KM99] G. Kemper and G. Malle, Invariant fields of finite irreducible reflection groups. Math.
Ann. 315 (1999), 569–586.
[Ko89] V. Kolyvagin, On the Mordell-Weil and Shafarevich-Tate groups for Weil elliptic
curves, Math. USSR-Izv. 33 (1989), 473–499.
[Kr24] W. Krull, Die verschiedenen Artender Hauptidealringe. Sitzungsberichte der Heidel-
berg Akademie (1924) no. 6.
[Kr31] W. Krull, Allgemeine Bewertungstheorie. J. Reine Angew. Math. 167 (1931), 160–196.
[Kr37] W. Krull, Beiträge zur Arithmetik kommutativer Integritätsbereiche, III, zum Dimen-
sionsbegriff der Idealtheorie, Mat. Zeit 42 (1937), 745–766.
[Kr51] W. Krull, Jacobsonsche Ringe, Hilbertscher Nullstellensatz Dimensionen theorie.
Math. Z. 54 (1951), 354–387.
[Ku99] M. Kurihara, On the ideal class groups of the maximal real subfields of number fields
with all roots of unity. J. Eur. Math. Soc. (JEMS) 1 (1999), 35–49.
BIBLIOGRAPHY 383

[La87] D. Laksov, Radicals and Hilbert Nullstellensatz for not necessarily algebraically closed
fields. Enseign. Math. (2) 33 (1987), 323–338.
[La99] T.Y. Lam, Lectures on modules and rings. Graduate Texts in Mathematics, 189.
Springer-Verlag, New York, 1999.
[La05] T.Y. Lam, Introduction to quadratic forms over fields. Graduate Studies in Mathe-
matics, 67. American Mathematical Society, Providence, RI, 2005.
[La06] T.Y. Lam, Serre’s problem on projective modules. Springer Monographs in Mathe-
matics. Springer-Verlag, Berlin, 2006.
[LR08] T.Y. Lam and M.L. Reyes, A prime ideal principle in commutative algebra. J. Algebra
319 (2008), 3006–3027.
[Lg53] S. Lang, The theory of real places. Ann. of Math. (2) 57 (1953), 378–391.
[Lg02] S. Lang, Algebra. Revised third edition. Graduate Texts in Mathematics, 211.
Springer-Verlag, New York, 2002.
[LM] M.D. Larsen and P.J. McCarthy, Multiplicative theory of ideals. Pure and Applied
Mathematics, Vol. 43. Academic Press, New York-London, 1971.
[La05] E. Lasker, Zur Theorie der Moduln und Ideale. Math. Ann. 60 (1905), 19–116.
[La64] D. Lazard, Sur les modules plats. C. R. Acad. Sci. Paris 258 (1964), 6313–6316.
[Le72] C.R. Leedham-Green, The class group of Dedekind domains. Trans. Amer. Math. Soc.
163 (1972), 493–500.
[Le43] F.W. Levi, Contributions to the theory of ordered groups. Proc. Indian Acad. Sci.,
Sect. A. 17 (1943), 199–201.
[Le67] L. Lesieur, Divers aspects de la théorie des idéaux d’un anneau commutatif. Enseigne-
ment Math. 13 (1967), 75–87.
[Li33] F.A Lindemann, The Unique Factorization of a Positive Integer. Quart. J. Math. 4,
319-320, 1933.
[Ma48] A.I. Malcev, On the embedding of group algebras in division algebras (Russian), Dokl.
Akad. Nauk. SSSR 60 (1948), 1499–1501.
[Ma58] H.B. Mann, On integral bases. Proc. Amer. Math. Soc. 9 (1958), 167–172.
[Ma71] C.F. Martin, Unique factorization of arithmetic functions. Aequationes Math. 7
(1971), 211.
[M] H. Matsumura, Commutative ring theory. Translated from the Japanese by M. Reid.
Second edition. Cambridge Studies in Advanced Mathematics, 8. Cambridge Univer-
sity Press, Cambridge, 1989.
[Ma44] K. Matusita, Über ein bewertungstheoretisches Axiomensystem für die Dedekind-
Noethersche Idealtheorie. Jap. J. Math. 19 (1944), 97–110.
[McC76] J. McCabe, A Note on Zariski’s Lemma. Amer. Math. Monthly 83 (1976), 560–561.
[Mi] J.W. Milnor, Topology from the differentiable viewpoint. Based on notes by David W.
Weaver. The University Press of Virginia, Charlottesville, Va. 1965.
[Mo49] T. Motzkin, The Euclidean algorithm. Bull. Amer. Math. Soc. 55 (1949), 1142–1146.
[Na57] M. Nagata, A remark on the unique factorization theorem. J. Math. Soc. Japan 9
(1957), 143–145.
[Na68] M. Nagata, A type of subrings of a noetherian ring. J. Math. Kyoto Univ. 8 (1968),
465–467.
[Na78] M. Nagata, On Euclid algorithm. C. P. Ramanujam-a tribute, pp. 175–186, Tata Inst.
Fund. Res. Studies in Math., 8, Springer, Berlin-New York, 1978.
[Na85] M. Nagata, Some remarks on Euclid rings. J. Math. Kyoto Univ. 25 (1985), 421–422.
[Na05] A.R. Naghipour, A simple proof of Cohen’s theorem. Amer. Math. Monthly 112 (2005),
825–826.
[Na53] N. Nakano, Idealtheorie in einem speziellen unendlichen algebraischen Zahlkörper. J.
Sci. Hiroshima Univ. Ser. A. 16: 425–439.
[Na95] W. Narkiewicz, A Note on Elasticity of Factorizations. J. Number Theory 51 (1995),
46–47.
[Ne49] B.H. Neumann, On ordered division rings. Trans. Amer. Math. Soc. 66 (1949), 202–
252.
[Ne07] M.D. Neusel, Invariant theory. Student Mathematical Library, 36. American Mathe-
matical Society, Providence, RI, 2007.
[No21] E. Noether, Idealtheorie in Ringberereichen. Math. Ann. 83 (1921), 24–66.
384 BIBLIOGRAPHY

[No26] E. Noether, Der Endlichkeitsatz der Invarianten endlicher linearer Gruppen der
Charakteristik p. Nachr. Ges. Wiss. G-ttingen: 28-35.
[NT] P.L. Clark, Number Theory: A Contemporary Introduction.
http://www.math.uga.edu/∼pete/4400FULL.pdf
[O74] T. Ogoma, On a problem of Fossum. Proc. Japan Acad. 50 (1974), 266–267.
[Ol69a] J.E. Olson, A combinatorial problem on finite Abelian groups. I. J. Number Theory
1 (1969), 8–10.
[Ol69b] J.E. Olson, A combinatorial problem on finite Abelian groups. II. J. Number Theory
1 (1969), 195–199.
[Pa59] Z. Papp, On algebraically closed modules. Publ. Math. Debrecen 6 (1959), 311–327.
[Pe04] H. Perdry, An elementary proof of Krull’s intersection theorem. Amer. Math. Monthly
111 (2004), 356–357.
[P] A. Pfister, Quadratic forms with applications to algebraic geometry and topology.
London Mathematical Society Lecture Note Series, 217. Cambridge University Press,
Cambridge, 1995.
[Qu96] C.S. Queen, Factorial domains. Proc. Amer. Math. Soc. 124 (1996), no. 1, 11–16.
[Qu76] D. Quillen, Projective modules over polynomial rings. Invent. Math. 36 (1976), 167–
171.
[Ra30] J.L. Rabinowitsch, Zum Hilbertschen Nullstellensatz. Math. Ann. 102 (1930), 520.
[R] M. Reid, Undergraduate commutative algebra. London Mathematical Society Student
Texts, 29. Cambridge University Press, Cambridge, 1995.
[Ro67] A. Robinson, Non-standard theory of Dedekind rings. Nederl. Akad. Wetensch. Proc.
Ser. A 70=Indag. Math. 29 (1967), 444–452.
[Ro93] M. Roitman, Polynomial extensions of atomic domains. J. Pure Appl. Algebra 87
(1993), 187–199.
[Ro76] M. Rosen, Elliptic curves and Dedekind domains. Proc. Amer. Math. Soc. 57 (1976),
197–201.
[Ro] J.J. Rotman, An introduction to homological algebra. Second edition. Universitext.
Springer, New York, 2009.
[Rü33] W. Rückert, Zum Eliminationsproblem der Potenzreihenideale, Math. Ann. 107
(1933), 259–281.
[Ru87] W. Rudin, Real and complex analysis. Third edition. McGraw-Hill Book Co., New
York, 1987.
[Sa61] P. Samuel, On unique factorization domains. Illinois J. Math. 5 (1961), 1–17.
[Sa64] P. Samuel, Lectures on unique factorization domains. Notes by M. Pavaman Murthy.
Tata Institute of Fundamental Research Lectures on Mathematics, No. 30 Tata Insti-
tute of Fundamental Research, Bombay 1964.
[Sa68] P. Samuel, Unique factorization. Amer. Math. Monthly 75 (1968), 945–952.
[Sa71] P. Samuel, About Euclidean rings. J. Algebra 19 (1971), 282–301.
[Sa08] A. Sasane, On the Krull dimension of rings of transfer functions. Acta Appl. Math.
103 (2008), 161- -168.
[S] W. Scharlau, Quadratic and Hermitian forms. Grundlehren der Mathematischen Wis-
senschaften 270. Springer-Verlag, Berlin, 1985.
[Sc45] T. Schönemann, Grundzüge einer allgemeinen Theorie der höhern Congruenzen,
deren Modul eine reelle Primzahl ist. J. Reine Angew. Math. 31 (1845), 269–325.
[Sc46] T. Schönemann, Von denjenigen Moduln, welche Potenzen von Primzahlen sind. J.
Reine Angew. Math. 32 (1846), 93–105.
[Sc46b] O.F.G. Schilling, Ideal theory on open Riemann surfaces. Bull. Amer. Math. Soc. 52
(1946), 945–963.
[Se56] A. Seidenberg, Some remarks on Hilbert’s Nullstellensatz. Arch. Math. (Basel) 7
(1956), 235–240.
[Si86] J. Silverman, The Arithmetic of Elliptic Curves, Graduate Texts in Mathematics 106,
Springer Verlag, 1986.
[ST42] A.H. Stone and J.W. Tukey, Generalized “sandwich” theorems. Duke Math. J. 9
(1942), 356–359.
[Su11] B. Sury, Uncountably Generated Ideals of Functions. College Math. Journal 42 (2011),
404–406.
BIBLIOGRAPHY 385

[Su76] A.A. Suslin, Projective modules over polynomial rings are free. Dokl. Akad. Nauk
SSSR 229 (1976), 1063–1066.
[Sw62] R.G. Swan, Vector bundles and projective modules. Trans. Amer. Math. Soc. 105
(1962), 264–277.
[Sw67] R.G. Swan, The number of generators of a module. Math. Z. 102 (4) (1967) 318-–322.
[Sw69] R.G. Swan, Invariant rational functions and a problem of Steenrod. Invent. Math. 7
(1969), 148–158.
[Te66] G. Terjanian, Sur les corps finis. C. R. Acad. Sci. Paris Sér. A-B 262 (1966), A167–
A169.
[Te72] G. Terjanian, Dimension arithmétique d’un corps. J. Algebra 22 (1972), 517–545.
[Tr88] H.F. Trotter, An overlooked example of nonunique factorization. Amer. Math.
Monthly 95 (1988), no. 4, 339–342.
[Va90] R.J. Valenza, Elasticity of factorizations in number fields. J. Number Theory 36
(1990), 212–218.
[vdW39] B.L. van der Waerden, Einftihrung in die algebraische Geometrie, Berlin, 1939.
[Wa] R.B. Warfield, Jr., Rings whose modules have nice decompositions. Math. Z. 125 (1972)
187–192.
[We07] J.H.M. Wedderburn, On Hypercomplex Numbers. Proc. of the London Math. Soc. 6
(1907), 77-118.
[W] C.A. Weibel, An introduction to homological algebra. Cambridge Studies in Advanced
Mathematics, 38. Cambridge University Press, Cambridge, 1994.
[We80] R.O. Wells Jr., Differential analysis on complex manifolds. Second edition. Graduate
Texts in Mathematics, 65. Springer-Verlag, New York-Berlin, 1980.
[Za04] F. Zanello, When are there infinitely many irreducible elements in a principal ideal
domain? Amer. Math. Monthly 111 (2004), 150–152.
[Za76] A. Zaks, Half factorial domains. Bull. Amer. Math. Soc. 82 (1976), 721–723.
[Za47] O. Zariski, A new proof of Hilbert’s Nullstellensatz. Bull. Amer. Math. Soc. 53 (1947),
362–368.
[Za69] H. Zassenhaus, On Hensel factorization. I. J. Number Theory 1 (1969), 291–311.
[Ze34] E. Zermelo, Elementare Betrachtungen zur Theorie der Primzahlen. Nachr. Gesellsch.
Wissensch. G-ttingen 1, 43-46, 1934.

You might also like